Remote Measurement of Turbulent Wind Spectra by Heterodyne DopplerLidar Technique

Philippe Drobinski Laboratoire de Météorologie Dynamique, Ecole Polytechnique, Palaiseau, France

Search for other papers by Philippe Drobinski in
Current site
Google Scholar
PubMed
Close
,
Alain M. Dabas Météo-France, Centre National de Recherches Météorologiques, Toulouse, France

Search for other papers by Alain M. Dabas in
Current site
Google Scholar
PubMed
Close
, and
Pierre H. Flamant Laboratoire de Météorologie Dynamique, Ecole Polytechnique, Palaiseau, France

Search for other papers by Pierre H. Flamant in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Heterodyne Doppler lidars (HDLs) are used to monitor atmospheric wind field and wind turbulence at remote distance. This last application calls for the derivation of wind spectra, which can be characterized by the dissipation rate and the κ-spectral peak (or outer scale of turbulence). However, the HDL technique may suffer two problems. First, HDL measurements result in spatial averaging of the true wind velocity along the line of sight, because of the laser pulse duration and windowing effect on processed signals. Second, even at high signal-to-noise ratio, the retrieved turbulent velocity field may be contaminated by errors due to speckle fluctuations. It is shown that both spatial averaging and error contribution to the wind spectra can be modeled starting from the transmitted laser pulse characteristics and signal processing parameters, so that their effect can be predicted. The rms difference between the estimated and predicted turbulent spectra is minimized in order to infer the turbulence parameters. This procedure is tested on simulated signals and validated on actual data taken by a 10-μm HDL during a field campaign in 1995.

The data collected during two periods of two consecutive days (9 and 10 March and 13 and 14 March 1995) are analyzed. On these days, moderate to light winds prevailed. The stability parameter zi/LMO indicated slightly unstable conditions with sometimes probable convection. The HDL measured energy dissipation rates ranging between 0.7 × 10−3 and 8 × 10−3 m2 s−3 in good agreement with sonic anemometer measurements. The κ-spectral peak ranged between 200 and 600 m.

Corresponding author address: Dr. Philippe Drobinski, Laboratoire de Meteorologie Dynamique, Ecole Polytechnique, 91128 Palaiseau Cedex, France.

Abstract

Heterodyne Doppler lidars (HDLs) are used to monitor atmospheric wind field and wind turbulence at remote distance. This last application calls for the derivation of wind spectra, which can be characterized by the dissipation rate and the κ-spectral peak (or outer scale of turbulence). However, the HDL technique may suffer two problems. First, HDL measurements result in spatial averaging of the true wind velocity along the line of sight, because of the laser pulse duration and windowing effect on processed signals. Second, even at high signal-to-noise ratio, the retrieved turbulent velocity field may be contaminated by errors due to speckle fluctuations. It is shown that both spatial averaging and error contribution to the wind spectra can be modeled starting from the transmitted laser pulse characteristics and signal processing parameters, so that their effect can be predicted. The rms difference between the estimated and predicted turbulent spectra is minimized in order to infer the turbulence parameters. This procedure is tested on simulated signals and validated on actual data taken by a 10-μm HDL during a field campaign in 1995.

The data collected during two periods of two consecutive days (9 and 10 March and 13 and 14 March 1995) are analyzed. On these days, moderate to light winds prevailed. The stability parameter zi/LMO indicated slightly unstable conditions with sometimes probable convection. The HDL measured energy dissipation rates ranging between 0.7 × 10−3 and 8 × 10−3 m2 s−3 in good agreement with sonic anemometer measurements. The κ-spectral peak ranged between 200 and 600 m.

Corresponding author address: Dr. Philippe Drobinski, Laboratoire de Meteorologie Dynamique, Ecole Polytechnique, 91128 Palaiseau Cedex, France.

Introduction

Preliminary measurements of wind turbulence in the planetary boundary layer (PBL) have been made with heterodyne Doppler lidars (HDLs) in the late 1980s and early 1990s. For instance, Ancellet et al. (1989) report an experiment where the dissipation rate of the turbulent kinetic energy (TKE) is estimated from the spectral broadening of HDL signals. Eberhard et al. (1989) derived vertical profiles of momentum fluxes in the PBL from the radial velocities measured by an HDL that scanned the laser beam conically around the vertical axis. Similar observations based on perpendicular range-height scans are also reported by Gal-Chen et al. (1992). In the same paper, the concept of estimating the dissipation rate of TKE from wind spectra is introduced. In this method, the time series of the vertical velocity acquired at a given range from a vertically pointing HDL is Fourier transformed. The resulting spectrum displays the expected −5/3 slope from which the dissipation rate may be estimated. Mayor et al. (1997) Fourier transformed range-resolved velocity profiles to recover the spectral behavior of the spatial fluctuations of the wind directly. No time-to-range conversion is needed. Frehlich et al. (1998) use another approach based on the estimation of the structure function of the radial velocity. They retrieve vertical profiles of TKE dissipation rate and outer length scale. In contrast to the previous works referenced previoulsy, Frehlich et al. (1998) account for the pulse averaging effect and measurement errors.

As far as the use of radars in turbulence measurement is concerned, several techniques have been developed and tested (Cohn 1995). They suffer a major limitation because of the wide divergence of the sounding beam and sidelobes. As compared with radars, lidars transmit “pencil” beams with small divergence, less than one milliradian. The size of the turbulence structures (or eddies) that are effectively sensed by active remote sensor, lidar or radar is a key factor. Lidars retrieve the strength of turbulence from the signature of turbulent eddies longer than the probing pulse (typically between 30 and 300 m), in contrast to radars that mostly rely on the spectral broadening brought by the intrapulse wind fluctuations.

The present article presents a new approach for measuring the dissipation rate and the κ-spectral peak (outer scale of TKE). In our study, the turbulence parameters are retrieved from wind spectra. Power spectra of HDL radial velocity profiles are computed and averaged over some time (that is, several profiles). A prediction of this spectrum based on Kolomogorov–Obukhov’s theory and including a proper treatment of pulse averaging and measurement errors is then fitted (in a least square sense) to the average spectrum computed from HDL velocities. The free parameters of the predictive model that achieve the best fit produce estimates for the dissipative rate and the outer scale of TKE. This method allows fast computation of turbulence quantities but is not suited to profiling because long-range intervals are required for the computation of wind spectra. Our approach presents the advantages of the HDL signal accumulation technique (Rye and Hardesty 1993) for the retrieval of the turbulence parameters and stresses the limitation of this technique in presence of turbulent eddy advection. A study of uncertainty on our spectral model leads to a condition for a reliable retrieval of wind spectra.

The theory and a discussion of a mathematical model accounting for pulse averaging effect and the contribution of measurement errors to velocity spectra is presented in section 2. The model is used in sections 3 and 4 to predict the turbulent spectrum of the radial velocities measured by the lidar. The prediction is tested on simulated HDL data. Then the possibility to estimate the TKE dissipation rate and the κ-spectral peak is introduced. It is tested on a set of actual HDL data acquired during the Boundary Layer Study in the Paris Area [Étude de la Couche Limite en Agglomération Parisienne (ECLAP)] experiment (Dupont et al. 1999). The turbulence parameters are qualitatively compared with the turbulence parameters deduced from a sonic anemometer deployed on the experiment site during the field campaign.

Lidar measurement

Pulse-averaging

The wind velocity υr(r) (see list of symbol in appendix A) measurement by HDL can be written as the sum of an effective wind velocity υm(r) and an error e(r),
υrrυmrer
According to Frehlich et al. (1998), the effective wind velocity υm(r) is a spatial average over the pulse duration and the processing range gate of the true line-of-sight (LOS) wind velocity υt(r),
i1520-0450-39-12-2434-E2
where the weighting function W(r) is given by
i1520-0450-39-12-2434-E3
The parameter G(r) is the power shape of the probing pulse [∫ G(r) dr = 1 for normalization] and Δp is the length of the processing range gate.
In the spectral domain, Eqs. (1) and (2) may be written as
i1520-0450-39-12-2434-E4
The quantity κ is the wavenumber, Φt(κ), Φm(κ), Φr(κ), Φe(κ), and ΦW(κ) are the power spectra of υt(r), υm(r), υr(r), e(r), and W(r), respectively (i.e., the squared magnitude of their Fourier transform).

Equation (4) assumes that e(r) and υm(r) have no cross correlation. This assumption calls for a justification since it is known that the measurement errors depend statistically on υm(r). For instance, the variance of errors increases with strong windshears or strong turbulent fluctuations. The statistical dependence does not imply, however, that there is a significant cross correlation between both quantities. Numerical studies (e.g., Frehlich et al. 1998) support the assumption that the cross-correlation term is negligible in the cloud-free atmosphere.

Equation (2) was proposed by Frehlich (1997) and Frehlich et al. (1998) for 2-μm HDL with Gaussian-shaped pulses with no chirp. The applicability to any HDL (not necessarily Gaussian shaped and linearly chirped) must be considered further.

Equation (2) is tested on simulated signals. The signals are generated by the “feuilleté” (“sliced”) model of Salamitou et al. (1995). The feuilleté model is an end-to-end numerical model that produces realistic HDL signals and may be used to assess the performance of HDL under nonstationary conditions. It combines laser beam propagation, interaction with the atmosphere, optical mixing, quadratic detection, and Doppler frequency estimators. The model slices the scattering medium along the beam axis, with each slice retaining the full transverse geometry of the problem. In such a feuilleté model, the individual particles or targets are first grouped within each slice and then a summation over the wavefields originating from the various slices is performed to determine the HDL signal. Because each slice backscatters an optical wavefield with a random phase delay, the summation displays speckle fluctuations. These fluctuations are independent from one model run to another. The input parameters are the complex amplitude of the laser pulse, the system efficiency accounting for the degrading effect of refractive index turbulence, a “true” velocity profile υt(r), and profiles for backscatter coefficient and atmospheric attenuation. Drobinski et al. (1999) have shown that at 10 μm, the system efficiency is degraded for C2n ≥ 10−14 m−2/3.

For the present tests, two different types of laser pulses are considered (Fig. 1). One is representative of a 10-μm HDL (Fig. 1, left panels). The laser pulse lasts several microseconds. It starts with a gain-switched-spike followed by a tail. It contains a quadratic frequency chirp (Willets and Harris 1982). The parameters used for the spike, the pulse duration, and frequency chirp are those of the experimental 10-μm HDL of the Laboratoire de Météorologie Dynamique (Delville et al. 1995). This system is described in section 3. The other laser pulse is representative of a 2-μm HDL (right panels of Fig. 1). The pulse shape is Gaussian, and the frequency chirp varies linearly with time. The parameters are taken from Frehlich et al. (1994).

The numerical experiment is based on the simulation of two sets of 500 signals (one set for each HDL). The backscatter, attenuation, and heterodyne system efficiencies are such that the average power of the signals is constant with range. The signals have all been obtained from the same input velocity profile υt(r) drawn from a Kolmogorov–Obukhov (Kolmogorov 1941; Obukhov 1941) spectrum
i1520-0450-39-12-2434-E6
where α = 2/3(2π)−2/3 (with C ≃ 0.76CK, where CK ≃ 2 is the Kolmogorov constant), ϵ is the dissipation rate of TKE, κin is the inner scale (κ−1in is the length of the smallest turbulent eddies, i.e., dissipation scale), and κout the κ-spectral peak or “plateau” (κ−1out is the size of the largest eddies). The quantity κ is in cycles per meter. The procedure used for generating υt from Φt is explained in appendix B. The range increment between the discrete velocity samples is δr = 1.5 m corresponding to a sampling frequency Fs = 100 MHz. For the present set of simulations, ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1 (the range increment δr is 1.5 m; it is the largest unaliased wavenumber), and κout = 0.002 m−1 (i.e., 500 m). The standard deviation συt of υt(r) is then 1 m s−1. In the present experiment, the signal-to-noise ratio (SNR) (ratio of signal power to noise power) is infinite. An infinite SNR has been retained here because it reduces the level of statistical uncertainty.

For each set, υr(r) is estimated by the pulse-pair frequency estimator (Miller and Rochwarger 1972). The 500 signals of each set are accumulated to retrieve a single profile. The frequencies are estimated over a sliding range gate of length Δp = 96 m. The number of signal samples per estimate is M = Δp/δr = 64. Considering the large number of accumulations, the precision for υr(r) is expected to be ≃0.10 m s−1 for the 10-μm HDL, and ≃0.02 m s−1 for the 2-μm HDL. Here, υm(r) is computed from υt(r) by using Eq. (2).

The comparison between υr(r) and υm(r) is presented in Fig. 2. The left and right panels are for the 10-μm and 2-μm HDL, respectively [the υt(r) profiles for 10-μm and 2-μm HDL are different because they result from two realizations of the turbulent wind field]. On both panels, υr(r) is drawn with a solid line, and υm(r) with a dashed line. The agreement between the curves is good, the rms of the differences (0.12 and 0.02 m s−1) are of the same order as the expected precision. Equation (2) is thus validated to the extended case of non-Gaussian and/or chirped laser pulses.

The low-pass filtering effect of pulse averaging is illustrated in Fig. 3 where the transfer function ΦW(κ) is displayed for the two different HDL pulses (see Fig. 1). A processing range gate Δp = 150 m is assumed. The figure displays a sharp transition between weakly and strongly attenuated wavenumbers at about κW = 3.0 × 10−3 m−1 and κW = 4.5 × 10−3 m−1 for the 10-μm and 2-μm HDL, respectively (κW is defined as the wavenumber achieving the attenuation of −10 dB). The figure sets the minimum size of turbulent eddies that can be detected by the HDL. In the present case, the sizes are 333 m and 222 m for the 10-μm and 2-μm HDL, respectively. Smaller scale eddies are filtered out. Figure 4 gives κ−1W as a function of Δp for a 10-μm HDL and a 2-μm HDL. For instance, it appears that for a 10-μm HDL a mimimum detectable eddy size of 500 m requires Δp ≃ 300 m, while for a 2-μm HDL, Δp ≃ 370 m.

Errors

The contribution of measurement errors to wind spectra derived from HDL data is a critical issue. In Frehlich et al. (1998) and Mayor et al. (1997), the error contribution is estimated from actual velocity profiles recorded by the HDL, assuming statistical independence of the errors between consecutive velocity profiles.

In this paper, the error contribution is modeled a priori, that is, related to HDL parameters. Presently, the model is only applicable to the pulse-pair frequency estimator. The major advantage is that the error contribution can be predicted prior to any measurements, based on a small number of parameters. Thus it makes possible the determination of a set of processing parameters that will guarantee negligible errors for specified conditions.

Error model

The error model is based on a set of equations that approximate the autocorrelation function (ACF) of errors. The error contribution to wind spectra is then determined by a Fourier transform [see Eq. (23) below]. The ACF of errors is calculated from the ACF of the range-resolved frequency estimates. The velocity υr(rl) and frequency f(l), both measured at a range rl = lδr are linked by
i1520-0450-39-12-2434-E7
where λ is the laser wavelength. Under the assumption that the statistical mean of e(rl) is zero (see next paragraph), we then obtain
i1520-0450-39-12-2434-E8
where the overbar denotes the statistical average, and, by substraction
i1520-0450-39-12-2434-E9
where f′(l) = f(l) − f(l). Finally,
i1520-0450-39-12-2434-E10
where the angle brackets 〈 · · · 〉 indicate statistical average.

The assumption that the errors have a zero mean is not true in reality unless several precautions are taken. For the pulse–pair frequency estimator reasons for the occurence of systematic errors are colored noise, occurence of gross errors at low SNR (Frehlich and Yadlowsky 1994), and a possible asymmetry of the power spectrum of the sounding pulse (Dabas et al. 1998). A colored noise can be avoided in practice by a proper matching of the analog antialiasing filter and the sampling frequency at detection level or by applying a digital whitening filter. Gross errors can also be avoided by restricting the processing of the signals to those with high SNR. As for the errors introduced by the asymetry of the pulse, it can be corrected (Dabas et al. 1998; Dabas et al. 2000). In the following, we assume that no systematic errors occur.

Let us now concentrate on the ACF of frequency estimates. Denoted by sn(p), the complex value of HDL signal n at range rp = pδr, the frequency f(l) estimated by the pulse–pair frequency estimator over the range interval [rl, rl+M−1] is given by
i1520-0450-39-12-2434-E11
with
i1520-0450-39-12-2434-E12
Here, Γ(l) is an estimate of the first-lag of the signal ACF averaged over the nacc accumulated signals. The superscript “*” denotes the complex conjugate.
A first-order expansion of Eq. (11) then provides with an approximate relationship between the random fluctuations of f(l) and Γ(l). Denoting Re and Im the real and imaginary parts of complex numbers, we have
i1520-0450-39-12-2434-E13
The correlation between the frequencies is derived in appendix C:
i1520-0450-39-12-2434-E14

The first- and second-order moments of f(l) can also be estimated from acquired velocity profiles (Mayor et al. 1997; similarly for water vapor by Kiemle et al. 1997), but they cannot be related to the Γ fluctuations by an analytical equation.

Equation (14) shows that the spatial autocorrelation of the frequency estimates depends on HDL signal statistics that cannot be related to HDL parameters unless some simplifying assumptions are made. In the following, we make the usual assumption of Gaussian HDL signal statistics and stationary processes. The optical wave captured by the HDL results from the addition of a great number of backscattered wavelets, which actually validates the Gaussian property. Stationarity, however, is an ideal property that HDL signals may approach over limited range and time intervals. It is nevertheless a reasonable approximation.

Under these assumptions, all fourth-order moments of the HDL signals can be split into a sum of second-order moments, and the statistics of HDL signals is described by the autocorrelation function of signal n, γn(p, q) = 〈s*n(p)sn(q)〉. The first-order moments in Eq. (14) can then be related to the second-order moments of the HDL signal
i1520-0450-39-12-2434-E15
The second-order moments are given by
i1520-0450-39-12-2434-E16

The ACF of measurement errors is now related to the single function γn(p, q). The next step is to determine γn(p, q). Two different approaches are possible: 1) estimating γn(p, q) from the signals, or 2) relating γn(p, q) to HDL and atmospheric parameters. The statistical estimate approach requires a large number of signals acquired under stationary conditions. In practice, this is possible with high repetition rate and stable HDL. We propose here to use the second approach.

A detailed study of the autocorrelation of HDL signals can be found in Frehlich (1994), and Frehlich and Cornman (1999). In both publications, equations are provided to relate the autocorrelation to instrument and atmospheric parameters. Assuming 1) that the atmospheric return (the useful part of the signal) and the contaminating noise are statistically independent, 2) that the contaminating noise is white (this assumption is justified above), and 3) that the optical and dynamical properties of the atmosphere are independent, then
i1520-0450-39-12-2434-E18
where δ(pq) is a Dirac and R(p, q) is the autocorrelation of the atmospheric return given by Frehlich and Cornman (1999):
i1520-0450-39-12-2434-E19
The quantity H(r) is the optical responsivity of the atmosphere, and υ′(r) the deviation of the true radial velocity from its linear average inside the probed volume. Equation (19) can be computed numerically assuming that H(r) is known and that wind fluctuations are modeled using Kolmogorov’s theory (e.g., Frehlich et al. 1998; Frehlich and Cornman 1999). However, to simplify the Eq. (19), 〈H(r)〉 can be assumed to be nearly constant with z (which is valid for a measurement inside an optically homogeneous volume), and 〈υ′2t(r)〉/ = σ2υt (Frehlich and Cornman 1999). Then the autocorrelation of the HDL signal becomes
i1520-0450-39-12-2434-E20
with
i1520-0450-39-12-2434-E21

Equation (20) offers the advantage that it can be conveniently applied since only the shape of the laser pulse and the variance of the radial velocity is required. Equation (20) can be used as a lower bound for the covariance, since σ2υt is larger than the mean-square velocity fluctuations over the range gate used in Frehlich and Cornman (1999). For a 10-μm HDL, the exponential term accounting for the decorrelation by turbulence can be neglected in most conditions, high turbulence levels are required to make it significant (ϵ ≥ 10−2 m2 s−3) (Giez et al. 1999; Dabas et al. 2000). This is at least true for the pulse–pair estimator because only the first lag of the ACF is used (|qp| = 1). If larger lags are used, the exponential term may impact on the decorrelation and the turbulence contribution may become important. For a 2-μm HDL, turbulence-induced decorrelation cannot be neglected and wind fluctuations must be taken into account using Eqs. (18) and (19).

As previously mentioned, the ACF of frequency estimates can be used to determine the spectral density of measurement errors. A good estimate for Φe is
i1520-0450-39-12-2434-E22
where nυ is the number of error profiles en over which the estimate is averaged, and L sets the length of the range gate over which the Fourier transform is performed. The mathematical expectation
i1520-0450-39-12-2434-E23
is then a good (i.e., nonbiased) prediction for Φe.

Validation

The autocorrelation function and the spectrum of measurement errors can be predicted from the laser pulse and the variance of the true radial wind using Eqs. (13), (16), (17), (20), and (23). We examine the relevance and practical significance of that prediction in this section.

The validation is based on processing simulated HDL signals. The simulation is conducted as in section 2.a. Two sets of 500 signals corresponding to the two different HDL shown in Fig. 1, are generated using a single velocity profile υt(r) corresponding to ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1 (i.e., 500 m). The standard deviation συt is 1 m s−1.

The simulated signals display independent speckle statistics and no noise. The frequency analysis is conducted with the pulse–pair frequency estimator. It is applied over a sliding range gate of length Δp. The number of accumulation nacc is varied. For each nacc, nυ = 500/nacc radial velocity profiles are produced. Whenever necessary, the profiles are afterward debiased (Dabas et al. 1998). The effective profiles υm(r) are computed from Eq. (2) and afterward substracted from each profile υr(r), which finally results in nυ independent error profiles en(r).

The first set of tests considers the prediction of the standard deviation of measurement errors. The prediction is obtained from Eq. (14) with l = d. The decorrelating effect of wind variability is taken into account with συt = 1 m s−1 in Eq. (20). The comparison between predicted and estimated standard deviations is shown in Figs. 5 and 6. In Fig. 5, the standard deviations for 10-μm and 2-μm HDL are displayed as a function of Δp for nacc = 1 and 5. The solid lines show predicted standard deviation of measurement errors, and the symbols “+” and “○,” the estimates. Good agreement is obtained. The main differences appear for short Δp and small nacc, that is, for large error levels. In that case, the first-order expansion Eq. (13) fails to produce an accuracte approximation of f(l).

Figure 6 extends our validation to low SNR for the particular case of a 10-μm HDL. The main source of noise for HDL is detection shot noise controlled by the local oscillator. Thermal and preamplifier noise are other sources of noise (Oh et al. 1996). The SNR is 0 dB. Compared to Fig. 5, the discrepancies between predicted and estimated standard deviations of measurement errors are larger and again the largest differences appear for short Δp and/or small nacc. Once again, we suspect the first-order expansion Eq. (13) is responsible. A practical limit for an accurate prediction of the error variance is the accumulation number and the SNR must be large enough.

The predictability of error spectra is investigated in Figs. 7 and 8 in which Φ̂e(κ) (solid lines) is compared with Φ̃e(κ) (dashed lines). Note that the predicted spectra Φ̃e(κ) are normalized so that they have the same variance as Φ̂e(κ). This normalization makes the comparison of the spectral shapes easier. The SNR is infinite in both cases and L = 8192. The spectral shapes of the errors are well predicted.

In addition, numerical investigations of the error spectrum have also been conducted for finite SNR that confirm the capability of our model to predict the shape of the spectral densities of errors.

As a consequence, the main limitation of our model turns out to be the variance of errors. In practice, the validity of the model is restricted to large accumulations and SNR as discussed below in section 3b.

Measurement of turbulence parameters

Section 2 presents a model that enables us to predict the error spectrum Φ̃e(κ) and so, assuming the errors of measurements are truly dominated by the random fluctuations of the signal (speckle and noise fluctuations), a prediction Φ̃r(κ) of Φr(κ) is given by
Φ̃r(κ)WκtκΦ̃e(κ)
The HDL characteristics provide ΦW(κ), γg, and SNR; Φt(κ) provides συt. These characteristics give an estimate of the signal ACF γ [see Eq. (20)] and Eqs. (14), (16), (17) give the ACF of the velocity measurements. Equation (23) transforms this latter function into a predicted error spectrum Φ̃e(κ), and at last, into a prediction of Φr(κ) through Eq. (24).

Inversion of this prediction allows us to estimate Φt(κ) from actual HDL data. We explore this possibility below.

Derivation of wind spectra

Given a set of nυ HDL profiles, an estimate for Φr(κ) is
i1520-0450-39-12-2434-E25
According to Eqs. (4) and (5), we should then have
Φ̂r(κ)Φ̂t(κ)WκΦ̃e(κ)
Below, we test Eq. (26) against the spectra derived from simulated HDL signals.

Figure 9 presents the results. It is based on the simulation of nυ = 50 sets of 10 signals generated from 50 different input velocity profiles υt(r), all drawn from the same Kolmogorov–Obukhov spectrum [see Eq. (6) with ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1]. The simulations consider a 10-μm HDL (left panels in Fig. 1). White noise is added so that SNR = 0 dB. Each set of 10 signals is accumulated (nacc = 10) to produce a single-velocity profile υr(r) (Δp = 100 m). Then, Φ̂r(κ) is computed according to Eq. (25) with L = 8192.

Figure 9 shows the contributions of Φm(κ) and Φ̃e(κ) to Φ̂r(κ). The original spectrum Φt(κ) is drawn with a dotted line. The attenuation due to pulse averaging is shown with Φm(κ) (dashed line) [see Eq. (5)]. The dash–dotted line is Φ̃e(κ). It shows that the contribution of the errors dominates above κ ≈ 2 × 10−3 m−1. In this region, the good agreement with Φ̂r(κ) displayed with a light solid line demonstrates the overall level of errors is well predicted. The heavy solid line is Φm(κ) + Φ̃e(κ). It agrees very well with Φ̂r(κ), which shows that the contributions of the real wind and the errors to the experimental spectrum Φ̂r(κ) are both well predicted by our model.

Statistical uncertainty

We study here the potential impact of statistical uncertainties on the “inversion” of Φ̂r(κ). Our aim is the determination of experimental conditions that guarantee a reliable inversion.

Denoting δΦ̂r(κ) = Φ̂r(κ) − Φr(κ) and δΦ̃e(κ) = Φ̃e(κ) − Φe(κ), we have
Φ̂m(κ)Φ̂r(κ)Φ̃e(κ)rκeκδΦ̂r(κδΦ̃e(κ)mκδΦ̂r(κδΦ̃e(κ)
So, an error δΦ̂m(κ) = δΦ̂r(κ)δΦ̃e(κ) is made on Φm(κ).

Since Φ̂r(κ) is the average of nυ spectra, each obtained from the Fourier transform of a single-velocity profile υr(r), it follows a chi-square distribution with nυ degrees of freedom [we assume υr(r) is Gaussian]. Its statistical average is 〈Φ̂r(κ)〉 = Φr(κ), and its standard deviation is n−1/2υΦr(κ).

In comparison, δΦ̃e(κ) is not random because both Φ̃e(κ) and Φe(κ) are deterministic quantities [Φ̃e(κ) is determined by a set of analytic equations, and Φe(κ) is a statistical average]. Its statistical average is δΦ̃e(κ), and its variance is zero.

The statistical average of δΦ̂m(κ) is then
δΦ̂m(κ)δΦ̃e(κ)
The error in the error model thus biases the estimate of Φm(κ).
Considering now the second-order moments, we obtain
i1520-0450-39-12-2434-E29
where σΦ̂m(κ), and σΦ̂r(κ) are the standard deviations of Φ̂m(κ), and Φ̂r(κ), respectively. In the following, we will consider that the estimate Φ̂m(κ) is accurate provided σΦ̂m(κ) ≪ Φm(κ). Thus the condition required for a useful measurement becomes
i1520-0450-39-12-2434-E30
Since Φe(κ) scales with n−1acc [this can be seen from Eqs. (16) and (17)], we can write
i1520-0450-39-12-2434-E31
where Φ1e(κ) is the “true” error spectrum when nacc = 1. Providing nυ ≫ 1, Eq. (30) becomes
i1520-0450-39-12-2434-E32

In Fig. 9, this condition is fulfilled up to κ = 3.5 × 10−3 m−1 (i.e., 286 m). In Fig. 10, we show that an accurate measurement Φ̂m(κ) of Φm(κ) can be made up to that wavenumber. The spectrum Φm(κ) is plotted with a dotted line and the solid line shows Φ̂m(κ) = Φ̂r(κ)Φ̃e(κ) over the frequency range 0 ⩽ κ ⩽ 3.5 × 10−3 m−1. The agreement is good. Note that the random fluctuations of Φ̂m(κ) become larger as the wavenumber increases. The reason is that, in the ratio Φe(κ)/Φm(κ), Φm(κ) tends to zero because of the pulse-averaging while Φe(κ) remains strictly positive. This results in the gradual saturation of condition Eq. (30).

Equation (32) sets a minimum value for naccn1/2υ needed to achieve an accurate retrieval of the wind spectrum on a spectral bandwidth [0; κW] for a given level of turbulence set by ϵ. Figure 11 displays the minimum value for naccn1/2υ to obtain a reliable spectrum in the frequency range [0, κW] for a 10-μm and 2-μm HDL, respectively. Figure 11 has been obtained by computing Φ1e from Eq. (23), assuming the error model is accurate, and Φm(κ) from Eq. (5). The maximum value of Φ1e(κ)/Φm(κ) over [0; κW] is determined and sets the minimum value for naccnυ according to Eq. (32). The quantity naccnυ is given as a function of Δp and dissipation rate ϵ [which partially characterizes Φm(κ)], considering no noise in HDL signals. Figure 11 can be used to set the minimum number of velocity profiles and signals to be accumulated in order to make an accurate measurement for a specified level of turbulence. Let us consider for instance ϵ = 10−4 m2 s−3. For a 10-μm HDL with Δp = 200 m (κ−1W ≃ 375 m), at least nυ ≈ 120 for nacc = 20 is required; that is, an acquisition time of 4 min at a 10-Hz pulse repetition frequency (PRF). For a 2-μm HDL, and Δp = 200 m (κ−1W ≃ 375 m), Eq. (32) gives nυ = 1 for nacc = 20. But, because the equation has been established on the condition that nυ ≫ 1, nυ ≈ 10 is preferable, so an acquisition time of 2 s is required for a 10-Hz PRF.

Limitations of the accumulation technique

In the previous sections, the simulations were conducted assuming an accumulation over nacc pulses probing the same velocity profile υt(r). This assumption may be not realistic, since υt(r) may vary from shot to shot due to wind advection (the spatial correlation follows Kolmogorov’s theory). Thus, an accumulation over nacc realizations filters out the turbulent wind fluctuations for measurements taken during the accumulation time Tacc = nacc/PRF. Taylor’s (1938) hypothesis views the turbulence as “frozen” in a field moving with a mean streamwise speed U. To satisfy the requirements that the turbulent eddies have negligible change as they advect, Willis and Deardorff (1976) suggest that
σUU
where σU is the standard deviation of the wind speed, and U is the mean wind speed. In this respect, the relevant streamwise wavenumber is κacc = (TaccU)−1. Figure 12a displays the spectral densities Φ̂r(κ) (thin line) and Φr(κ) for κout = 0.002 m−1 (i.e., 500 m) and ϵ = 7 × 10−3 m2 s−3 for a 10-μm HDL with Δp = 150 m. Between two shots, the velocity profile υt(r) drifts streamwise by 15 m (e.g., U = 15 m s−1 and PRF = 1 Hz. Because nacc = 50 HDL shots are accumulated, the turbulent fluctuations of the wind are smoothed over an equivalent length scale of κ−1acc = 750 m. In Fig. 12b, in which U = 3 m s−1 and PRF = 1 Hz, the impact of accumulation is negligible since κaccκW. In fact, the velocity profile υt(r) drifts only 3 m, yielding a spatial filtering of κ−1acc = 150 m with nacc = 50.
Since the contribution of HDL shot accumulation may have a large impact on the retrieval of the turbulent wind spectra, two constraints are now required for a relevant measurement. The first constraint sets the condition on the level of measurement errors with respect to the random fluctuations of wind profile υm(r). Equation (32) sets the minimum value naccn1/2υ to meet the measurement specifications. The second constraint refers to the number of HDL signals nacc that must be accumulated, provided that the accumulation period is short enough, that is κacc must be large enough to satisfy κWκacc. In summary, the conditions required for a reliable measurement are
i1520-0450-39-12-2434-E34

Let us consider two HDL operating at 10 and 2 μm, respectively. From Fig. 4, we have κ−1W are at best ≃300 m and ≃100 m, respectively. Table 1 gives the maximum number nacc of HDL shots that must be accumulated in order to satisfy condition 2 [see Eq. (34)] for various PRF and various values for U. Condition 1 sets the minimum number nυ of spectral densities to average as a function of expected turbulence. The present state of art in HDL technology allows high enough PRF that condition 2 is easily met.

Retrieval of turbulence parameters during the ECLAP experiment

The operational procedure presented above has been applied to actual signals using the 10-μm HDL operated by Laboratoire de Météorologie Dynamique (LMD) to retrieve κout and ϵ that characterize a fully developed turbulent state. The data were obtained during a field campaign (ECLAP) conducted during winter in 1994/ 95 to study planetary boundary layer (PBL) dynamics and meteorological conditions that could result in heavy pollution episodes (Dupont et al. 1999). During ECLAP, numerous in situ sensors including a sonic anemometer, were deployed on a site at the École Polytechnique located 25 km south of Paris.

The LMD 10-μm HDL provided wind radial velocity measurements along the LOS. The characteristics of the transmitted pulse are displayed in Fig. 1. Several time series of 200 shots at 2 Hz PRF (i.e., 100-s measurements) looking horizontally in the PBL were performed every ≃25 min at z ≃ 250 m (there is a slight elevation angle ≃0.8° along the LOS and, as shown in Fig. 13a, the terrain is approximately homogeneous between 2.8 and 7 km where SNR ≥ 0 dB, see Fig. 13b). Sonic anemometers at 10 and 30 m above the ground recorded wind and temperature measurements at a rate of 10 Hz with a precision of ≃1% (Hall et al. 1984; Carissimo and Gilbert 1992). These data were averaged over 10 min. The data were collected during two intensive observation periods (IOP), 9–10 and 13–14 March 1995, respectively (hereinafter called M09, M10, M13, and M14). Table 2 gives the mean and turbulent meteorological parameters and the PBL top, for the IOP. On these days, moderate to light winds prevailed. The stability parameter zi/LMO indicated slightly unstable conditions (zi/LMO ranges between −7.3 and −1.1, except for M13 in late afternoon where zi/LMO = 3.6) with sometimes probable convection as referred in Drobinski et al. (1998). Here, LMO is the Monin–Obukhov length, it is used to define the stability of the atmosphere (when LMO > 0, the PBL is stably stratified, when LMO = 0 it is neutrally stratified and when LMO < 0, it is unstably stratified). It is given by
i1520-0450-39-12-2434-E35
zi is the PBL inversion height determined by a 0.53-μm backscatter lidar pointing to zenith, and operated on the same site during the HDL measurements (Menut et al. 1999). The accuracy on zi is ≃30 m. The quantity k = 0.4 is the von Kármán constant, u∗ is the friction velocity and is related to the surface momentum flux, w∗ is the convective velocity scale and is related to surface buoyancy flux. The turbulence quantities u∗ and w∗ were computed from sonic anemometer measurements. The TKE dissipation rate at 10- and 30-m height is computed using the equation ϵ = u3[φM(z/LMO) − z/LMO]/kz and is displayed in Fig. 16. The quantity φM is the nondimensional wind shear. In stable stratification, LMO ≥ 0 and φM(z/LMO) = 1 + 4.7z/LMO. In unstable stratification, LMO ⩽ 0 and φM(z/LMO) = (1 − 15z/LMO)−/14 (Businger et al. 1971; Wyngaard and Coté 1971). The HDL radial velocity variance is approximately constant along the LOS where SNR ≥ 0 dB.

The operational procedure described in sections 2 and 3 is applied to the actual 10-μm HDL signals to retrieve κout and ϵ. The dissipation rates ϵ obtained by HDL at ≃250 m are compared (at least for their order of magnitude) to those from the sonic anemometer at 30-m height.

Retrieval of wind spectra

The turbulent wind spectra are estimated with an accumulation factor of nacc = 20 and by averaging the power spectrum of nυ = 10 velocity profiles (naccn1/2υ ≃ 64). According to Eq. (33) and Table 2, Taylor’s hypothesis is satisfied. The average wind speed ranged between 1 and 7 m s−1, with peak values larger than 10 m s−1. For nacc = 20, condition (2) is always fulfilled. The HDL was looking nearly horizontally in the PBL (elevation angle 0.8°). The SNR is better than 0 dB up to ranges of several kilometers. Thus, we expect that our error model is valid (negligible detection noise). The angle between the streamwise motion of turbulence and the HDL LOS was corrected before the retrieval of wind spectra (Banakh et al. 1995).

Figure 14 shows the spectral density Φ̂r(κ) (thin line) computed from one set of actual data. The spectral density is computed from the radial velocity profiles over the entire range (2.8 km; 7 km) in Fig. 14a, over (2.8 km; 5 km) range in Fig. 14b and over (5 km; 7 km) range in Fig. 14c to assess the natural variability of the turbulence parameters. The signal frequency is estimated over a sliding window of length Δp = 100 m (M = 17 signal samples per estimate with a sampling frequency Fs = 25 MHz. The mean of the nυ = 20 velocity profiles is removed before each profile is individually Fourier transformed. The Fourier transform is performed over the range interval 2.8–7 km (700 velocity estimates), where SNR ≥ 0 dB (Fig. 13b). The predicted spectrum Φ̃r(κ) closest to Φ̂r(κ) is displayed with a thick line. It has been derived by determining the turbulence quantities ϵ and κout of the Kolmogorov–Obukhov spectrum Φt(κ) that minimize the cost function
i1520-0450-39-12-2434-E36
For this example, the minimization over the entire range is reached for ϵ = 4.6 × 10−3 m2 s−3 and κout = 0.0022 m−1. The minimization over the subranges (2.8 km; 5 km) and (5 km; 7 km) is reached for ϵ = 3.7 × 10−3 m2 s−3 and κout = 0.0021 m−1, and ϵ = 5.2 × 10−3 m2 s−3 and κout = 0.0023 m−1, respectively. Figure 14 shows this minimization produces a good agreement between the predicted and the observed spectrum, at least in the spectral domain concerned by the minimization (κκW ≃ 0.004 m−1) (in Figs. 14b and c, the smaller number of samples may degrade the precision of the measurement). It can be shown that the contribution of errors is negligible in that particular domain. As a matter of fact, the application of the error-filtering technique from Mayor et al. (1997) to our velocity profiles results in a power spectrum very close to Φ̂r(κ) in the spectral interval κκW, confirming that the contribution of errors is negligible in that particular domain. Sensitivity studies using numerical simulations have shown that the precision on ϵ and κout estimates is ≃10% for κout ⩽ 0.004 m−1. The procedure used to derive the “instrument” precision was the following: 1) generation of simulated HDL signals (the total length is 8192 samples at 100 MHz sampling frequency, i.e., 12.3 km range, and 2 Hz PRF) accounting for a turbulent wind profile for SNR = +∞; 2) estimation of the radial velocity profile using the pulse–pair frequency estimator (with M = 64) from nacc = 20 accumulated HDL signals (i.e., 10-s accumulation) simulated with same υt(r) radial velocity profiles; 3) estimation of Φ̂r(κ) from nυ = 10 radial velocity profile estimates (so one spectrum used to compute ϵ and κout is estimated every 100 s); and 4) minimization of the cost function J(ϵ, κout), retrieval of the turbulence parameters ϵ and κout, and calculation of the departure with respect to the input values. This procedure was applied 10 000 times in order to have good statistics on the accuracy of the model for both ϵ and κout retrieval. The statistics are thus derived from 10 000 × nυ random realizations of the radial velocity profile as well as new speckle signals for each lidar shot. This precision accounts for the uncertainties arising from random fluctuations of the true wind field and of the HDL signals (speckle and noise). In the simulation, the impact of shot-to-shot pulse instabilities was not treated, the transmitted pulses are the same for all simulated signals and the nacc realizations used to process one velocity profile are uncorrelated. These assumptions have an impact on the calculation of the instrument error, and the value of 10%–20% must be considered as an optimistic bound. The variability of the turbulence parameters shown on Fig. 14 (≃15%–30%) is partly due to instrument statistical uncertainty (≃10%) and partly to natural variability (≃10%–30% assuming decorrelation between instrument precision and natural variability). The atmospheric variability is the dominant source of error in the following.

This value of κout corresponds to a length scale that is larger than the height above the boundary (i.e., ≃250 m), and to eddies that transport momentum. Hence, the corresponding eddies are anisotropic. The −5/3 slope of the radial velocity component’s κ spectra is not sufficient evidence for the existence of an “inertial subrange” that depends on the flow being locally isotropic (Kolmogorov 1941). Considering the altitude of the LOS above the surface (see Fig. 13a) or the κ-spectral peak determined from our analysis (see Fig. 15), these structures are likely to be outside the inertial subrange. That the −5/3 regime predicted by Kolmogorov’s theory extends above the inertial subrange has been reported in literature (Busch and Panofsky 1968; Vinnichenko and Dutton 1969; Donelan and Miyake 1973).

The reasons why our model in Fig. 14 underestimates the actual spectrum in the high-wavenumber domain (κκW) has not been determined precisely. However, it has no impact on the estimation of the turbulence quantities because it is outside the range over which the optimization (36) was performed. A possible explaination is pulse-to-pulse instability of the laser that makes the average frequency, the chirp and the shape of pulse fluctuate randomly from one shot to the other. The magnitude of the measurement errors may also be increased by natural wind variability and probably mainly by wind shear within the range gate. According to Dabas et al. (2000), the range of estimated ϵ values in the ECLAP experiment, using the LMD 10-μm HDL measurements, should not induce significative decorrelation due to turbulence, but Eq. (20) does not account for wind shear which may not be negligible.

Figure 15 is the histogram of κ−1out for the entire dataset. Two modes appear, one with a Gaussian shape around ≃410 m, and another one between 1 and 1.4 km. This second mode probably reflects the instability of the minimization technique in presence of organized convection [here roll vortices, see Etling and Brown (1993) for details]. The scale of such structures developed in these conditions is typically above 1 km (Drobinski et al. 1998). Table 2 shows that most of the HDL data were collected during slightly unstably stratified PBL when such structures are likely. During M13, periodicities appear in the data recorded by the sonic anemometer causing an approximate 20% variation. Strong peak in the energy appears on M13 with a 1.3-h period suggesting organized wave motion in the PBL flow. On M13 and M14, HDL radial velocity measurements show energy peaks in the spectral analysis corresponding to longitudinal roll wavelengths of 1.4 km (±0.2 km) and 1.6 km (±0.2 km), respectively (Drobinski et al. 1998). On M09 and M10, no organized wave motion was apparent. In the following, the organized large eddies are “removed” by smoothing the spectra, in order to stabilize the minimization.

The mean value for κ-spectral peak is κout ≃ 410 m with a standard deviation of ≃100 m, which is partly due to the model accuracy and partly due to natural variability (see Fig. 14). The mean value of κoutz is ≃0.66 (±0.2). Previous studies have shown that κoutz depends on stratification (Busch and Panofsky 1968; Van Mieghen 1973; Hojstrup 1982): 0.01 ⩽ κoutz ⩽ 0.3 for free convection, 0.3 ⩽ κoutz ⩽ 1 for forced convection and 0.6 ⩽ κoutz in case of isotropic turbulence. Considering the present values of κoutz, they are consistent with forced convection and isotropic turbulence regimes.

Measurement of the TKE dissipation rate

The measurements of TKE dissipation rate by HDL are compared with values computed from the turbulent surface fluxes measured by a sonic anemometer at the same time. The surface measurements of energy dissipation rate ϵ are approximately corrected for altitude dependence, using (Hojstrup 1982; Kristensen et al. 1989; Frehlich et al. 1998):
i1520-0450-39-12-2434-E37
Hojstrup in 1982 concluded that his model [Eq. (37)] provided a good description of the PBL turbulence up to a height of 0.5zi under stability conditions ranging from neutral throughout the unstable range, over homegeneous terrain. Our HDL measurements were taken in these conditions with z/zi ⩽ 0.4 (except on M14 where z/zi ≃ 0.6–0.7). The model described by Eq. (37) is no more reliable when z/zi ≥ 0.5, for in the upper half of the PBL, velocity spectra are influenced by entrainment processes and effect of inversion (Hojstrup 1982). These effects were described qualitatively by Hojstrup (1982) and were not included in the model. Studies accounting for shear- and buoyancy-driven PBL flows were conducted since then using large eddy simulations (LES) (Moeng 1984; Moeng and Wyngaard 1988; Mason 1989; Moeng and Sullivan 1994). Moeng and Sullivan (1994) derived from their simulations an analytical model for TKE dissipation rate accounting for shear and buoyancy:
i1520-0450-39-12-2434-E38
Figure 17 displays the turbulence parameters from the sonic anemometer measurements on M09, M10, M13, and M14. The thin solid and dashed lines represent u∗ and w∗, respectively. The thick solid and dashed lines represent ϵ at 250-m height computed from Eq. (37) (Hojstrup 1982) and Eq. (38) (Moeng and Sullivan 1994), respectively. According to the relative errors on zi estimates and on sonic measurements (precision ≃1%), the instrument error bar for ϵ is ≃3%–9% (power of three of the turbulent fluxes). For the four days, the temporal variability of u∗ and w∗ is ≃20% of their mean values. These fluctuations induce a natural variability on the computed TKE dissipation rates of ≃30% of their mean values, showing that natural variability dominates over instrument error (Lenschow and Stankov 1986; Lenschow et al. 1994). The values of ϵ computed from Hojstrup or Moeng and Sullivan models are in good agreement (the difference between the two models is on average ⩽1.5 × 10−4 m2 s−3, or ≃6% of the ϵ mean value, that is, within the statistical uncertainty of our measurements).

Figure 18a compares lidar and sonic anemometer measured energy dissipation rates using Eq. (37) (represented with circles) and Eq. (38) (represented with stars). Numerical sensitivity study has shown that the model precision on ϵ estimates is ≃10% for κout ⩽ 0.004 m−1 (this values does not account for natural variability). The instrument statistical uncertainty is shown with vertical bars (±1 − σ). The comparison indicates both sensors obtain the same order of magnitude, which is encouraging considering the uncertainties of measurements. The rms error is 1.7 × 10−3 m2 s−3 and 1.4 × 10−3 m2 s−3 for Eqs. (37) and (38), respectively. It is due partly to the intrinsic errors in the measurements by the sonic anemometer and HDL, partly to errors introduced by the height correction, partly to the different geometries of the measurement (point measurements and spatial measurements), partly due to the spatial and temporal variability and, also to the signature of large-scale wind features.

The ratio of Eqs. (37) and (38) ranges between 1 and 2 at most. The difference between the two models does not affect the order of magnitude. In particular, the bias of 2 × 10−4 m2 s−3 for ϵ computed with Eq. (37) [8 × 10−4 m2 s−3 for ϵ computed with Eq. (38)] is not significant because, according to Fig. 11, naccn1/2υ = 64 allows a reliable measurement for ϵ ≥ 10−3 m2 s−3. This TKE dissipation rate measurement precision is thus not fully adapted to turbulence studies because of the large pulse length. Using an HDL with shorter pulse duration will improve this measurement.

The horizontal velocity standard deviation σ̂υt, calculated from the turbulent wind spectrum estimates Φ̂t(κ) [see Eq. (26)] (characterized by ϵ and κout) are compared with the best-fit spectral model (Panofsky et al. 1977; Kristensen et al. 1989). This model does not vary with height,
i1520-0450-39-12-2434-E39
The spatial integration of Φ̂t(κ) needed to derive σ̂υt is equivalent to a 10-min integration time for the sonic anemometer measurements. Figure 18b compares lidar and sonic anemometer measured horizontal velocity standard deviation. The instrument statistical uncertainty is shown with vertical bars (±1 − σ). The comparison gives the right order of magnitude. The bias is 0.11 m s−1, and the rms error is 0.10 m s−1. This comparison looks better than that for ϵ for this is a lower order term than ϵ [see Eqs. (37), (38) and (39), the instrument error bar for συ is ≃1%]. The results are consistent between the lidar and the sonic anemometer measurements and give confidence in the estimated values of ϵ and κout. Nevertheless, these are preliminary results and the precision of ϵ and κout measurements have to be assessed later using an HDL with shorter pulse length.

Conclusions

Turbulent wind spectra Φ̂r(κ) retrieved by heterodyne Doppler lidar (HDL) can be reliabily converted into spectra Φ̂t(κ) of point radial velocities provided a correction for spatial averaging of the velocity field is performed. This correction is required because of the finite duration of the HDL pulse and the contribution of measurement errors. The error model proposed in this paper is valid as long as the signal-to-noise ratio SNR ≥ 0 dB, otherwise it may need an accumulation of many HDL shots. The spatial average performed by the probing pulse truncates the wind turbulence spectra, as does an accumulation on several successive measurements.

Spectral analysis of the radial velocities, retrieved using the 10-μm HDL operated by Laboratoire de Météorologie Dynamique, during the ECLAP experiment produced a κ−5/3 power law. The data were collected during two IOPs on 9–10 and 13–14 March 1995, respectively. On these days, moderate to light winds prevailed, with the stability parameter zi/LMO, ranging between −7.3 and −1.1, indicating slightly unstable conditions sometimes with organized large eddies (Drobinski et al. 1998). Our estimate of turbulent energy dissipation rate ϵ using the HDL technique was consistent with the average computed from the sonic anemometer measurements. This is an indirect validation of our analysis. Corrections for height between the two instruments used two analytical models derived by Hojstrup (1982) and Moeng and Sullivan (1994). The energy dissipation rates measured by these two instruments ranged between 0.7 × 10−3 and 8 × 10−3 m2 s−3, which is moderate turbulence (McCready 1964). The κ-spectral peak ranged between 200 and 600 m.

Our method allows fast computation of turbulent wind spectra and retrieval of turbulence quantities. It is not suited to profiling. Comparatively, the estimation of a structure function can be made locally [see Eq. (39) in Frehlich et al. (1998)], and is suited to profiling. However, PBL wind fields may display organized structures such as organized roll vortices [their spacing is generally ≃1–5 km; see Etling and Brown (1993) for a review], streaks [their spacing is ≃100 m, and they are embedded in an ≃100-m layer depth, see, e.g., Moeng and Sullivan (1994); Weckwerth et al. (1996); Foster (1997)], or other coherent structures. These structures have a well-defined signature in the spectral domain that make them easily detectable, but their signature on a structure function may not be as straightforward. A detailed comparison of the spectral and time-domain (structure function) approaches is left to future works.

At present, a limitation of the HDL technique is due to limited range resolution, so these results must be considered as preliminary results. For planetary boundary layer studies, a desirable resolution is on the order (or below) of 100 m. A 2-μm HDL can achieve this resolution. Future works will also be directed to the possibility of improving the error model at lower SNR.

Acknowledgments

The authors would like to thank the anonymous referees that helped to improve the manuscript significantly; R. C. Foster and B. J. Rye for their comments and contribution to edit the final manuscript;M. Farge for fruitful discussions on turbulence theory;and P. Delville, B. Romand, C. Boitel, L. Menenger, and C. Loth for their assistance during the field campaign ECLAP. They are grateful to L. Menut and E. Dupont for providing the in situ sensors data. The work was conducted at the Laboratoire de Météorologie Dynamique du CNRS and Météo-France, it was supported by the European Space Agency (ESA), the Centre National d’Études Spatiales (CNES), and Alcatel (formerly Aérospatiale Cannes).

REFERENCES

  • Ancellet, G. M., R. T. Menzies, and W. B. Grant, 1989: Atmospheric velocity spectral width measurements using the statistical distribution of pulsed CO2 lidar return signal intensities. J. Atmos. Oceanic Technol.,6, 50–58.

  • Banakh, V. A., C. Werner, N. N. Kerkis, F. Köpp, and I. N. Smalikho, 1995: Turbulence measurements with a CW Doppler lidar in the atmospheric boundary layer. Atmos. Oceanic Opt.,8, 955–959.

  • Busch, N. E., and H. A. Panofsky, 1968: Recent spectra of atmospheric turbulence. Quart. J. Roy. Meteor. Soc.,94, 132–148.

  • Businger, J. A., J. C. Wyngaard, Y. Izumi, and E. F. Bradley, 1971: Flux profile relationships in the atmospheric surface layer. J. Atmos. Sci.,28, 181–189.

  • Carissimo, B., and E. Gilbert, 1992: Un anémomètre ultrasonique pour la mesure des paramètres moyens et turbulents de l’atmosphère. Tech. Note EDF HE33/95-02. [Available from B. Carissimo, EDF, Chatou, France].

  • Cohn, S. A., 1995: Radar measurements of turbulent eddy dissipation rate in the troposphere: A comparison of techniques. J. Atmos. Oceanic Technol.,12, 85–95.

  • Dabas, A. M., P. Drobinski, and P. H. Flamant, 1998: Chirp induced bias in velocity measurements by a coherent Doppler CO2 lidar. J. Atmos. Oceanic Technol.,15, 407–415.

  • ——, ——, and ——, 2000: Velocity biases of adaptive filter estimates in heterodyne Doppler lidar measurements. J. Atmos. Oceanic Technol.,17, 1189–1202.

  • Delville, P., C. Loth, P. H. Flamant, D. Bruneau, T. Le Floch, and J. C. Darcy, 1995: A new TE − CO2 laser for coherent lidar and wind applications. Proc. Eighth Coherent Laser Radar Conf., Keystone, CO, Optical Society of America, 297–300.

  • Donelan, M., and M. Miyake, 1973: Spectra and fluxes in the boundary layer of the trade-wind zone. J. Atmos. Sci.,30, 444–464.

  • Drobinski, P., R. A. Brown, P. H. Flamant, and J. Pelon, 1998: Evidence of organized large eddies by ground-based Doppler lidar, sonic anemometer and sodar. Bound.-Layer Meteor.,88, 343–361.

  • ——, A. M. Dabas, P. Delville, P. H. Flamant, J. Pelon, and R. M. Hardesty, 1999: Refractive-index structure parameter in the planetary boundary layer: Comparison of measurements taken by a 10.6-μm coherent lidar, a 0.9 μm scintillometer, and in situ sensors. Appl. Opt.,38, 1648–1656.

  • Dupont, E., L. Menut, B. Carissimo, J. Pelon, and P. H. Flamant, 1999: Comparison between the atmospheric boundary layer in Paris and its rural suburbs during the ECLAP experiment. Atmos. Environ.,33, 979–994.

  • Eberhard, W. L., R. E. Cupp, and K. R. Healy, 1989: Doppler lidar measurements of profiles of turbulence and momentum flux. J. Atmos. Oceanic Technol.,6, 809–819.

  • Etling, D., and R. A. Brown, 1993: Roll vortices in the planetary boundary layer: A review. Bound.-Layer Meteor.,65, 215–248.

  • Foster, R. C., 1997: Structure and energetics of optimal Ekman layer perturbations. J. Fluid. Mech.,333, 97–123.

  • Frehlich, R. G., 1994: Coherent Doppler lidar signal covariance including wind shear and wind variance. Appl. Opt.,33, 6472–6481.

  • ——, 1997: Effects of wind turbulence on coherent Doppler lidar performance. J. Atmos. Oceanic Technol.,14, 54–75.

  • ——, and M. J. Yadlowsky, 1994: Performance of mean-frequency estimators for Doppler radar and lidar. J. Atmos. Oceanic Technol.,11, 1217–1230.

  • ——, and L. B. Cornman, 1999: Coherent Doppler lidar signal spectrum with wind turbulence. Appl. Opt.,38, 7456–7466.

  • ——, S. M. Hannon, and S. W. Henderson, 1994: Performance of a 2-μm coherent Doppler lidar for wind measurements. J. Atmos. Oceanic Technol.,11, 1517–1528.

  • ——, ——, and ——, 1998: Coherent Doppler lidar measurements of wind field statistics. Bound.-Layer Meteor.,86, 233–256.

  • Gal-Chen, T., M. Xu, and W. L. Eberhard, 1992: Estimations of atmospheric boundary layer fluxes and other turbulence parameters from Doppler lidar data. J. Geophys. Res.,97, 18 409–18 423.

  • Giez, A., G. Ehret, R. L. Schwiesow, K. J. Davis, and D. H. Lenschow, 1999: Water vapor flux measurements from ground-based vertically pointed water vapor differential absorption and Doppler lidars. J. Atmos. Oceanic Technol.,16, 237–250.

  • Hall, F. F., Jr., R. M. Huffaker, R. M. Hardesty, M. E. Jackson, T. R. Lawrence, M. J. Post, R. A. Richter, and B. F. Weber, 1984: Wind measurement accuracy of NOAA pulsed infrared Doppler lidar. Appl. Opt.,23, 2503–2506.

  • Hojstrup, J., 1982: Velocity spectra in the unstable planetary boundary layer. J. Atmos. Sci.,39, 2239–2248.

  • Kiemle, C., G. Ehret, A. Giez, K. J. Davis, D. H. Lenschow, and S. P. Oncley, 1997: Estimation of boundary layer humidity fluxes and statistics from airborne differential absorption lidar (DIAL). J. Geophys. Res.,102, 29 189–29 203.

  • Kolmogorov, A. N., 1941: Energy dissipation in locally isotropic turbulence. Dokl. AN SSSR,32, 19–21.

  • Kristensen, L., D. H. Lenschow, P. Kirkegaard, and M. Courtney, 1989: The spectral velocity tensor for homogeneous boundary layer turbulence. Bound.-Layer Meteor.,47, 149–193.

  • Lenschow, D. H., and B. B. Stankov, 1986: Length scales in the convective boundary layer. J. Atmos. Sci.,43, 1198–1209.

  • ——, J. Mann, and L. Kristensen, 1994: How long is long enough when measuring fluxes and other turbulent statistics? J. Atmos. Oceanic Technol.,11, 661–673.

  • Mason, P. J., 1989: Large eddy simulation of the convective atmospheric boundary layer. J. Atmos. Sci.,46, 1492–1516.

  • Mayor, S. D., D. Lenshow, R. L. Schweisow, J. Mann, C. L. Frush, and M. K. Simon, 1997: Validation of NCAR 10.6-μm CO2 Doppler lidar radial velocity measurements and comparison with a 915-MHz profiler. J. Atmos. Oceanic Technol.,14, 1110–1126.

  • McCready, P. B., Jr., 1964: Standardization of gustiness values from aircraft. J. Appl. Meteor.,3, 439–449.

  • Menut, L., C. Flamant, J. Pelon, and P. H. Flamant, 1999: Urban boundary-layer height determination from lidar measurements over the Paris area. Appl. Opt.,38, 945–954.

  • Miller, K. S., and M. M. Rochwarger, 1972: A covariance approach to spectral moment estimation. IEEE Trans. Inform. Theory,IT-18, 588–596.

  • Moeng, C. H., 1984: A large eddy simulation model for the study of planetary boundary layer turbulence. J. Atmos. Sci.,41, 2052–2062.

  • ——, and J. C. Wyngaard, 1988: Spectral analysis of large-eddy simulations of the convective boundary layer. J. Atmos. Sci.,45, 3575–3587.

  • ——, and P. P. Sullivan, 1994: A comparison of shear- and buoyancy-driven planetary boundary layer flows. J. Atmos. Sci.,51, 999–1022.

  • Obukhov, A. M., 1941: Energy distribution in the spectrum of a turbulent flow. Izv. AN SSSR, Ser. Geogr. Geofiz.,5, 453–466.

  • Oh, D., P. Drobinski, P. Salamitou, and P. H. Flamant, 1996: Optimal local oscillator power for CMT photo-voltaic detector in heterodyne mode. Infrared Phys. Tech.,37, 325–333.

  • Panofsky, H. A., H. Tennekes, D. H. Lenschow, and J. C. Wyngaard, 1977: The characteristics of turbulent velocity components in the surface layer under convective conditions. Bound.-Layer Meteor.,11, 355–361.

  • 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, 16–27.

  • Salamitou, P., A. M. Dabas, and P. H. Flamant, 1995: Simulation in the time domain for heterodyne coherent laser radar. Appl. Opt.,34, 499–506.

  • Taylor, G. I., 1938: The spectrum of turbulence. Proc. Roy. Soc. London,A132, 476–490.

  • Van Mieghen, J., 1973: Atmospheric Energetics. Clarendon Press, 306 pp.

  • Vinnichenko, N. K., and J. A. Dutton, 1969: Empirical studies of atmospheric structure and spectra in the free atmosphere. Radio Sci.,4, 1115–1126.

  • Weckwerth, T. M., C. J. Grund, and S. D. Mayor, 1996: Linearly organized coherent structures in the surface layer. Preprints, Symp. on Boundary Layers and Turbulence, Vancouver, BC, Canada, Amer. Meteor. Soc., 22–23.

  • Willets, D. V., and M. R. Harris, 1982: An investigation into the origin of frequency sweeping in a hybrid TEA CO2 laser. J. Phys. D: Appl. Phys.,15, 51–67.

  • Willis, G. E., and J. W. Deardorff, 1976: On the use of Taylor’s translation hypothesis for diffusion in the mixed layer. Quart. J. Roy. Meteor. Soc.,102, 817–822.

  • Wyngaard, J. C., and O. R. Coté, 1971: The budgets of turbulent kinetic energy and temperature variance in the atmospheric surface layer. J. Atmos. Sci.,28, 190–201.

  • Zrnić, D. S., 1979: Estimation of spectral moments of weather echoes. IEEE Trans. Geosci. Electron.,GE-17, 113–128.

APPENDIX A

List of Symbols

  • CK   Kolmogorov constant

  • C2n   Refractive index structure parameter

  • c   Light velocity

  • E(κ)   Energy spectrum function

  • en(r)   nth profile of measurement errors

  • f   Frequency estimate

  • g(t)   Laser pulse shape

  • H(r)   Optical responsivity of the atmosphere

  • K   Instrumental constant

  • k   von Kármán constant

  • L   Overall signal length in range increments δr

  • LMO   Monin–Obukhov length

  • LOS   Line of sight

  • l   Index for L

  • M   Number of time samples in a processing range gate

  • m   Index for M

  • nacc   Number of accumulated HDL shots

  • nυ   Number of independent spectra

  • n   Index for nacc or nυ

  • PRF   Pulse repetition frequency

  • R   Autocorrelation of the atmospheric return

  • r, r′   Space variables

  • SNR   Broadband signal-to-noise ratio

  • s   HDL signal

  • Tacc   Accumulation time

  • Ts   Sampling period

  • t   Time variable

  • U   Wind velocity

  • u∗   Friction velocity

  • υm(r)   Theoretical turbulent wind velocity including spatial averaging

  • υr(r)   Theoretical turbulent wind velocity including spatial averaging and errors

  • υt(r)   True turbulent wind velocity

  • υt(r)   Deviation of the true turbulent wind velocity

  • W(r)   Range weighting function

  • w∗   Convective velocity scale

  • z   Vertical coordinate

  • zi   PBL inversion height

  • Δp   Processing range gate length

  • δ   Dirac function

  • δr   Range increment cTs/2

  • δij   Kronecker symbol

  • δΦ̃e(κ)   Error on Φe(κ)

  • δΦ̂m(κ)   Error on Φm(κ)

  • δΦ̂r(κ)   Error on Φr(κ)

  • ϵ   Turbulent energy dissipation rate

  • Γ   HDL signal autocorrelation function estimate

  • γ   HDL signal autocorrelation function

  • γg   Autocorrelation function of the laser pulse

  • γnoise   Autocorrelation function of the noise

  • κ   Wavenumber

  • κacc   Streamwise wavenumber such that κacc = (TaccU)−1

  • κin   Inner scale of turbulence

  • κout   Outer scale of turbulence, or κ-spectral peak, or “plateau”

  • κW   Cutoff wavenumber such that ΦW(κW) = 0.1

  • Φe(κ)   True measurement error spectrum

  • Φ̃e(κ)   Model for the measurement error spectrum

  • Φ̂e(κ)   Estimate of the measurement error spectrum

  • Φm(κ)   Theoretical spectral density of υm(r)

  • Φ̂m(κ)   Estimate of Φm(κ)

  • Φr(κ)   Theoretical spectral density of υr(r)

  • Φ̂r(κ)   Estimate of Φr(κ)

  • Φt(κ)   True spectral density

  • Φ̂t(κ)   Estimate of Φt(κ)

  • ΦW(κ)   Frequency response of W(z)

  • σU   Standard deviation of U

  • συ   Best-fit spectral model for velocity standard deviation

  • συt   Standard deviation of υt(r)

  • σ̂υt   Estimate of συt

  • σΦ̂m(κ)   Standard deviation of Φ̂m(κ)

  • σΦ̂r(κ)   Standard deviation of Φ̂r(κ)

APPENDIX B

Simulation of Turbulent Velocity Profiles

The appendix shows how a random realization of a turbulent velocity profile can be obtained from a power spectrum Φt(κ) such as the one expressed in Eq. (6).

The procedure starts with sampling Φt(κ) at the regularly spaced wavenumbers ki = (i/N − 0.5)/δr, with i = 0, . . . , N − 1, δr is the range increment for the generated velocity profile, and N is the length (in units of δr).

The random velocity profile is then obtained by the following inverse Fourier transform:
i1520-0450-39-12-2434-EB1
where k = 0, . . . , N − 1, and x(i) is a complex number normally distributed with the following properties:
i1520-0450-39-12-2434-EB2
with δi,j being the Kronecker symbol.

APPENDIX C

Derivation of Eqs. (13) and (14)

The first- and second-order statistics of (2πTs)f(l) = arg[Γ(l)] cannot be derived analytically unless it is assumed that the random fluctuations Γ′(l) of Γ(l) around its statistical average Γ(l) are small. Following mathematical derivations already presented by Zrnić (1979) for the particular case of the error variance, a first-order expansion of (2πTs)f(l) is
i1520-0450-39-12-2434-EC1
The subscripts Re and Im denote the real and imaginary parts, respectively. We then have
i1520-0450-39-12-2434-EC2
From Eq. (C2), it is now possible to derive an approximate expression for the correlation of the random fluctuations f′(l) of f(l) around its average f(l). Let us consider first the frequencies f(l) and f(d) computed from the lth and dth range gates (containing the signal samples l to l + M − 1, and d to d + M − 1, respectively). Using the properties on operations with complex numbers
i1520-0450-39-12-2434-EC3
it can be shown that
i1520-0450-39-12-2434-EC5

Fig. 1.
Fig. 1.

Typical (left) 10-μm and (right) 2-μm HDL pulses. Here (a) and (b) are the power shapes and (c) and (d) are the frequency chirps.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 2.
Fig. 2.

Comparison between the retreived υr(r) (solid line) and expected υm(t) (dashed line) velocity profiles, for (a) 10-μm and (b) 2-μm HDL for Δp = 96 m in both cases. The value υr(r) is estimated by pulse–pair frequency estimator. The 500 HDL signals are accumulated. Here, υm(r) is computed from Eq. (2); υt(r) is drawn from a Kolmogorov–Obukhov spectrum with ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 3.
Fig. 3.

Frequency response ΦW(κ) of the weighting function W(z). The cutoff wavenumber κW corresponds to ΦW(κW) = 0.1. Here, Δp = 150 m leading to κW = 3.0 × 10−3 m−1 (i.e., ≃330 m) and 4.5 × 10−3 m−1 (i.e., ≃220 m) for (a) 10-μm and (b) 2-μm HDL, respectively.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 4.
Fig. 4.

Here, κ−1W as function of Δp for (a) 10-μm and (b) 2-μm HDL, respectively.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 5.
Fig. 5.

Standard deviation of the measurement errors as a function of Δp, for (a) 10-μm and (b) 2-μm HDL, and no detection noise. Here, nacc = 1 and nacc = 5 are considered. For each one of them, the standard deviation predicted by the theory is displayed with solid lines, whereas the standard deviations computed from the simulations are shown with crosses (nacc = 1) and circles (nacc = 5). The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κout = 0.002 m−1, and κin = 0.33 m−1.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 6.
Fig. 6.

Same as Fig. 5 (10-μm HDL), but white noise is added to the signals (SNR = 0 dB). Three accumulation factors have been considered: nacc = 1 (crosses), nacc = 5, (circles) and nacc = 10 (triangles).

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 7.
Fig. 7.

Power spectra of the measurement errors for various signal processing configurations for a 10-μm HDL with SNR = +∞. The solid lines are the spectra derived from the simulations, the dashed lines are drawn from theory. Both experimental and theoretical spectra have been normalized to unit area to make the comparison easier. (a) Δp = 45 m and nacc = 1; (b) Δp = 45 m and nacc = 5; (c) Δp = 450 m and nacc = 1; (d) Δp = 450 m and nacc = 5. The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 8.
Fig. 8.

Same as Fig. 7, but for a 2-μm HDL.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 9.
Fig. 9.

Comparison between spectra retrieved from processed signals (thin solid line) and predicted by theory (heavy solid line). The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1. The retrieved spectrum has been obtained from 50 sets of 10 simulated signals generated from 50 independent velocity profiles υt(r). The laser wavelength is 10 μm. The processing range gate is Δp = 100 m. The velocity measurements are performed with the pulse–pair frequency estimator with accumulation of 10 signals, so 50 velocity profiles υr(r) are retrieved. The power spectrum of the initial velocity profiles υt(r) is displayed with a dotted line. The dashed line then includes the pulse averaging effect. The dash–dotted line shows the contribution of measurement errors.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 10.
Fig. 10.

Velocity spectrum Φm(κ) (dotted line) of the original radial velocity profiles υt(r), in comparison with Φ̂m(κ) (solid line) retrieved from the measurements. The turbulence characteristics of υt(r) and the processing parameters are the same as in Fig. 9.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 11.
Fig. 11.

Contour plot giving the minimum value for naccn1/2υ required to compute a reliable spectrum, as a function of Δp and ϵ, for (a) 10-μm and (b) 2-μm HDL. The turbulence parameters are κout = 0.002 m−1 and κin = 0.33 m−1.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 12.
Fig. 12.

Comparison between spectra retrieved from processed signals (thin solid line) and predicted by theory (heavy solid line) for a 10-μm HDL. Between two shots, the turbulent wind profile υt(r) is advected by (a) 15 m and by (b) 3 m, respectively. Here, nacc = 50 and nυ = 20. The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κout = 0.002 m−1, and κin = 0.33 m−1.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 13.
Fig. 13.

(a): HDL line of sight (dashed line) from the Laboratoire de Météorologie Dynamique in Palaiseau with respect to the underlying terrain (solid line). (b): SNR as a function of range. The statistical uncertainty on the whole dataset is shown with vertical bars (±1 − σ).

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 14.
Fig. 14.

Comparison between spectra retrieved from processed actual signals (thin solid line), recorded by 10-μm HDL operated by Laboratoire de Météorologie Dynamique (LMD) during the ECLAP experiment on M13 at 1430 UTC, and predicted by theory (heavy solid line). The turbulence parameters used to compute the theoretical spectrum Φr(κ) are (a) ϵ = 4.6 × 10−3 m2 s−3 and κout = 0.0022 m−1 on the entire range (2.8 km; 7 km), (b) ϵ = 3.7 × 10−3 m2 s−3 and κout = 0.0021 m−1 on (2.8 km; 5 km) range, and (c) ϵ = 5.2 × 10−3 m2 s−3 and κout = 0.0023 m−1 on (5 km; 7km) range. The corresponding turbulence parameters measured by the sonic anemometer and the backscatter lidar on M13 at 1430 UTC are u∗ = 0.67 m s−1, w∗ = 1.49 m s−1, zi = 0.87 km, LMO = −198 m, ϵ = 5.0 × 10−3 m2 s−3, and 3.0 × 10−3 m2 s−3 for Eqs. (37) and (38), respectively.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 15.
Fig. 15.

Histogram of outer scale of turbulence κ−1out retrieved by LMD 10-μm HDL during ECLAP experiment. The instrument precision on κout estimates is ≃10%–20%. The natural variability for κout is of the same order of magnitude (see Fig. 14).

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 16.
Fig. 16.

Time evolution of the TKE dissipation rate ϵ from sonic anemometer measurements at 10 (solid line) and 30 m (dashed line) height on (a) M09, (b) M10, (c) M13, and (d) M14. The TKE dissipation rate is computed from the relationship ϵ = u3[φM(z/LMO) − z/LMO]/kz.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 17.
Fig. 17.

Turbulent parameters from sonic anemometer measurements on (a) M09, (b) M10, (c) M13, and (d) M14. The thin solid and dashed lines represent u∗ and w∗, respectively. The thick solid and dashed lines represent ϵ computed from Eq. (37) (Hojstrup 1982) and Eq. (38) (Moeng and Sullivan 1994), respectively.

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Fig. 18.
Fig. 18.

(a) Scatterplot of best-fit spectral model for ϵ [Hojstrup (1992) model with circles and Moeng and Sullivan (1994) model with stars] computed from sonic anemometer measurements vs ϵ retrieved by LMD 10-μm HDL during ECLAP experiment. The instrument statistical uncertainty is shown with vertical bars (±1 − σ). (b) Scatterplot of best-fit spectral spectral model for the velocity standard deviation (Panofsky et al. 1977; Kristensen et al. 1989) computed from sonic anemometer measurements vs σ̂υt computed from Φ̂t(κ) estimated from the inversion procedure. The instrument statistical uncertainty is shown with vertical bars (±1 − σ).

Citation: Journal of Applied Meteorology 39, 12; 10.1175/1520-0450(2000)039<2434:RMOTWS>2.0.CO;2

Table 1.

Maximal number of accumulated lidar signals nacc fulfilling condition 2 (see text) for a 10- and 2-μm HDL with κ−1W of 300 and 100 m, respectively. Between parentheses, is the minimum number nv that validates condition 1 (see text) for κout = 0.002 m−1 (i.e., 500 m), ϵ = 7 × 10−3 m2 s−3, and Δp = 100 m.

Table 1.
Table 2.

Average mean and turbulence PBL parameters measured at Palaiseau on 9, 10, 13, and 14 Mar 1995 (hereon called M09, M10, M13, and M14). Here, zi is the PBL height, u is the friction velocity, w is the convective velocity scale, LMO is the Monin–Obukhov length, and σU is the standard deviation of the wind velocity U.

Table 2.
Save
  • Ancellet, G. M., R. T. Menzies, and W. B. Grant, 1989: Atmospheric velocity spectral width measurements using the statistical distribution of pulsed CO2 lidar return signal intensities. J. Atmos. Oceanic Technol.,6, 50–58.

  • Banakh, V. A., C. Werner, N. N. Kerkis, F. Köpp, and I. N. Smalikho, 1995: Turbulence measurements with a CW Doppler lidar in the atmospheric boundary layer. Atmos. Oceanic Opt.,8, 955–959.

  • Busch, N. E., and H. A. Panofsky, 1968: Recent spectra of atmospheric turbulence. Quart. J. Roy. Meteor. Soc.,94, 132–148.

  • Businger, J. A., J. C. Wyngaard, Y. Izumi, and E. F. Bradley, 1971: Flux profile relationships in the atmospheric surface layer. J. Atmos. Sci.,28, 181–189.

  • Carissimo, B., and E. Gilbert, 1992: Un anémomètre ultrasonique pour la mesure des paramètres moyens et turbulents de l’atmosphère. Tech. Note EDF HE33/95-02. [Available from B. Carissimo, EDF, Chatou, France].

  • Cohn, S. A., 1995: Radar measurements of turbulent eddy dissipation rate in the troposphere: A comparison of techniques. J. Atmos. Oceanic Technol.,12, 85–95.

  • Dabas, A. M., P. Drobinski, and P. H. Flamant, 1998: Chirp induced bias in velocity measurements by a coherent Doppler CO2 lidar. J. Atmos. Oceanic Technol.,15, 407–415.

  • ——, ——, and ——, 2000: Velocity biases of adaptive filter estimates in heterodyne Doppler lidar measurements. J. Atmos. Oceanic Technol.,17, 1189–1202.

  • Delville, P., C. Loth, P. H. Flamant, D. Bruneau, T. Le Floch, and J. C. Darcy, 1995: A new TE − CO2 laser for coherent lidar and wind applications. Proc. Eighth Coherent Laser Radar Conf., Keystone, CO, Optical Society of America, 297–300.

  • Donelan, M., and M. Miyake, 1973: Spectra and fluxes in the boundary layer of the trade-wind zone. J. Atmos. Sci.,30, 444–464.

  • Drobinski, P., R. A. Brown, P. H. Flamant, and J. Pelon, 1998: Evidence of organized large eddies by ground-based Doppler lidar, sonic anemometer and sodar. Bound.-Layer Meteor.,88, 343–361.

  • ——, A. M. Dabas, P. Delville, P. H. Flamant, J. Pelon, and R. M. Hardesty, 1999: Refractive-index structure parameter in the planetary boundary layer: Comparison of measurements taken by a 10.6-μm coherent lidar, a 0.9 μm scintillometer, and in situ sensors. Appl. Opt.,38, 1648–1656.

  • Dupont, E., L. Menut, B. Carissimo, J. Pelon, and P. H. Flamant, 1999: Comparison between the atmospheric boundary layer in Paris and its rural suburbs during the ECLAP experiment. Atmos. Environ.,33, 979–994.

  • Eberhard, W. L., R. E. Cupp, and K. R. Healy, 1989: Doppler lidar measurements of profiles of turbulence and momentum flux. J. Atmos. Oceanic Technol.,6, 809–819.

  • Etling, D., and R. A. Brown, 1993: Roll vortices in the planetary boundary layer: A review. Bound.-Layer Meteor.,65, 215–248.

  • Foster, R. C., 1997: Structure and energetics of optimal Ekman layer perturbations. J. Fluid. Mech.,333, 97–123.

  • Frehlich, R. G., 1994: Coherent Doppler lidar signal covariance including wind shear and wind variance. Appl. Opt.,33, 6472–6481.

  • ——, 1997: Effects of wind turbulence on coherent Doppler lidar performance. J. Atmos. Oceanic Technol.,14, 54–75.

  • ——, and M. J. Yadlowsky, 1994: Performance of mean-frequency estimators for Doppler radar and lidar. J. Atmos. Oceanic Technol.,11, 1217–1230.

  • ——, and L. B. Cornman, 1999: Coherent Doppler lidar signal spectrum with wind turbulence. Appl. Opt.,38, 7456–7466.

  • ——, S. M. Hannon, and S. W. Henderson, 1994: Performance of a 2-μm coherent Doppler lidar for wind measurements. J. Atmos. Oceanic Technol.,11, 1517–1528.

  • ——, ——, and ——, 1998: Coherent Doppler lidar measurements of wind field statistics. Bound.-Layer Meteor.,86, 233–256.

  • Gal-Chen, T., M. Xu, and W. L. Eberhard, 1992: Estimations of atmospheric boundary layer fluxes and other turbulence parameters from Doppler lidar data. J. Geophys. Res.,97, 18 409–18 423.

  • Giez, A., G. Ehret, R. L. Schwiesow, K. J. Davis, and D. H. Lenschow, 1999: Water vapor flux measurements from ground-based vertically pointed water vapor differential absorption and Doppler lidars. J. Atmos. Oceanic Technol.,16, 237–250.

  • Hall, F. F., Jr., R. M. Huffaker, R. M. Hardesty, M. E. Jackson, T. R. Lawrence, M. J. Post, R. A. Richter, and B. F. Weber, 1984: Wind measurement accuracy of NOAA pulsed infrared Doppler lidar. Appl. Opt.,23, 2503–2506.

  • Hojstrup, J., 1982: Velocity spectra in the unstable planetary boundary layer. J. Atmos. Sci.,39, 2239–2248.

  • Kiemle, C., G. Ehret, A. Giez, K. J. Davis, D. H. Lenschow, and S. P. Oncley, 1997: Estimation of boundary layer humidity fluxes and statistics from airborne differential absorption lidar (DIAL). J. Geophys. Res.,102, 29 189–29 203.

  • Kolmogorov, A. N., 1941: Energy dissipation in locally isotropic turbulence. Dokl. AN SSSR,32, 19–21.

  • Kristensen, L., D. H. Lenschow, P. Kirkegaard, and M. Courtney, 1989: The spectral velocity tensor for homogeneous boundary layer turbulence. Bound.-Layer Meteor.,47, 149–193.

  • Lenschow, D. H., and B. B. Stankov, 1986: Length scales in the convective boundary layer. J. Atmos. Sci.,43, 1198–1209.

  • ——, J. Mann, and L. Kristensen, 1994: How long is long enough when measuring fluxes and other turbulent statistics? J. Atmos. Oceanic Technol.,11, 661–673.

  • Mason, P. J., 1989: Large eddy simulation of the convective atmospheric boundary layer. J. Atmos. Sci.,46, 1492–1516.

  • Mayor, S. D., D. Lenshow, R. L. Schweisow, J. Mann, C. L. Frush, and M. K. Simon, 1997: Validation of NCAR 10.6-μm CO2 Doppler lidar radial velocity measurements and comparison with a 915-MHz profiler. J. Atmos. Oceanic Technol.,14, 1110–1126.

  • McCready, P. B., Jr., 1964: Standardization of gustiness values from aircraft. J. Appl. Meteor.,3, 439–449.

  • Menut, L., C. Flamant, J. Pelon, and P. H. Flamant, 1999: Urban boundary-layer height determination from lidar measurements over the Paris area. Appl. Opt.,38, 945–954.

  • Miller, K. S., and M. M. Rochwarger, 1972: A covariance approach to spectral moment estimation. IEEE Trans. Inform. Theory,IT-18, 588–596.

  • Moeng, C. H., 1984: A large eddy simulation model for the study of planetary boundary layer turbulence. J. Atmos. Sci.,41, 2052–2062.

  • ——, and J. C. Wyngaard, 1988: Spectral analysis of large-eddy simulations of the convective boundary layer. J. Atmos. Sci.,45, 3575–3587.

  • ——, and P. P. Sullivan, 1994: A comparison of shear- and buoyancy-driven planetary boundary layer flows. J. Atmos. Sci.,51, 999–1022.

  • Obukhov, A. M., 1941: Energy distribution in the spectrum of a turbulent flow. Izv. AN SSSR, Ser. Geogr. Geofiz.,5, 453–466.

  • Oh, D., P. Drobinski, P. Salamitou, and P. H. Flamant, 1996: Optimal local oscillator power for CMT photo-voltaic detector in heterodyne mode. Infrared Phys. Tech.,37, 325–333.

  • Panofsky, H. A., H. Tennekes, D. H. Lenschow, and J. C. Wyngaard, 1977: The characteristics of turbulent velocity components in the surface layer under convective conditions. Bound.-Layer Meteor.,11, 355–361.

  • 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, 16–27.

  • Salamitou, P., A. M. Dabas, and P. H. Flamant, 1995: Simulation in the time domain for heterodyne coherent laser radar. Appl. Opt.,34, 499–506.

  • Taylor, G. I., 1938: The spectrum of turbulence. Proc. Roy. Soc. London,A132, 476–490.

  • Van Mieghen, J., 1973: Atmospheric Energetics. Clarendon Press, 306 pp.

  • Vinnichenko, N. K., and J. A. Dutton, 1969: Empirical studies of atmospheric structure and spectra in the free atmosphere. Radio Sci.,4, 1115–1126.

  • Weckwerth, T. M., C. J. Grund, and S. D. Mayor, 1996: Linearly organized coherent structures in the surface layer. Preprints, Symp. on Boundary Layers and Turbulence, Vancouver, BC, Canada, Amer. Meteor. Soc., 22–23.

  • Willets, D. V., and M. R. Harris, 1982: An investigation into the origin of frequency sweeping in a hybrid TEA CO2 laser. J. Phys. D: Appl. Phys.,15, 51–67.

  • Willis, G. E., and J. W. Deardorff, 1976: On the use of Taylor’s translation hypothesis for diffusion in the mixed layer. Quart. J. Roy. Meteor. Soc.,102, 817–822.

  • Wyngaard, J. C., and O. R. Coté, 1971: The budgets of turbulent kinetic energy and temperature variance in the atmospheric surface layer. J. Atmos. Sci.,28, 190–201.

  • Zrnić, D. S., 1979: Estimation of spectral moments of weather echoes. IEEE Trans. Geosci. Electron.,GE-17, 113–128.

  • Fig. 1.

    Typical (left) 10-μm and (right) 2-μm HDL pulses. Here (a) and (b) are the power shapes and (c) and (d) are the frequency chirps.

  • Fig. 2.

    Comparison between the retreived υr(r) (solid line) and expected υm(t) (dashed line) velocity profiles, for (a) 10-μm and (b) 2-μm HDL for Δp = 96 m in both cases. The value υr(r) is estimated by pulse–pair frequency estimator. The 500 HDL signals are accumulated. Here, υm(r) is computed from Eq. (2); υt(r) is drawn from a Kolmogorov–Obukhov spectrum with ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1.

  • Fig. 3.

    Frequency response ΦW(κ) of the weighting function W(z). The cutoff wavenumber κW corresponds to ΦW(κW) = 0.1. Here, Δp = 150 m leading to κW = 3.0 × 10−3 m−1 (i.e., ≃330 m) and 4.5 × 10−3 m−1 (i.e., ≃220 m) for (a) 10-μm and (b) 2-μm HDL, respectively.

  • Fig. 4.

    Here, κ−1W as function of Δp for (a) 10-μm and (b) 2-μm HDL, respectively.

  • Fig. 5.

    Standard deviation of the measurement errors as a function of Δp, for (a) 10-μm and (b) 2-μm HDL, and no detection noise. Here, nacc = 1 and nacc = 5 are considered. For each one of them, the standard deviation predicted by the theory is displayed with solid lines, whereas the standard deviations computed from the simulations are shown with crosses (nacc = 1) and circles (nacc = 5). The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κout = 0.002 m−1, and κin = 0.33 m−1.

  • Fig. 6.

    Same as Fig. 5 (10-μm HDL), but white noise is added to the signals (SNR = 0 dB). Three accumulation factors have been considered: nacc = 1 (crosses), nacc = 5, (circles) and nacc = 10 (triangles).

  • Fig. 7.

    Power spectra of the measurement errors for various signal processing configurations for a 10-μm HDL with SNR = +∞. The solid lines are the spectra derived from the simulations, the dashed lines are drawn from theory. Both experimental and theoretical spectra have been normalized to unit area to make the comparison easier. (a) Δp = 45 m and nacc = 1; (b) Δp = 45 m and nacc = 5; (c) Δp = 450 m and nacc = 1; (d) Δp = 450 m and nacc = 5. The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1.

  • Fig. 8.

    Same as Fig. 7, but for a 2-μm HDL.

  • Fig. 9.

    Comparison between spectra retrieved from processed signals (thin solid line) and predicted by theory (heavy solid line). The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κin = 0.33 m−1, and κout = 0.002 m−1. The retrieved spectrum has been obtained from 50 sets of 10 simulated signals generated from 50 independent velocity profiles υt(r). The laser wavelength is 10 μm. The processing range gate is Δp = 100 m. The velocity measurements are performed with the pulse–pair frequency estimator with accumulation of 10 signals, so 50 velocity profiles υr(r) are retrieved. The power spectrum of the initial velocity profiles υt(r) is displayed with a dotted line. The dashed line then includes the pulse averaging effect. The dash–dotted line shows the contribution of measurement errors.

  • Fig. 10.

    Velocity spectrum Φm(κ) (dotted line) of the original radial velocity profiles υt(r), in comparison with Φ̂m(κ) (solid line) retrieved from the measurements. The turbulence characteristics of υt(r) and the processing parameters are the same as in Fig. 9.

  • Fig. 11.

    Contour plot giving the minimum value for naccn1/2υ required to compute a reliable spectrum, as a function of Δp and ϵ, for (a) 10-μm and (b) 2-μm HDL. The turbulence parameters are κout = 0.002 m−1 and κin = 0.33 m−1.

  • Fig. 12.

    Comparison between spectra retrieved from processed signals (thin solid line) and predicted by theory (heavy solid line) for a 10-μm HDL. Between two shots, the turbulent wind profile υt(r) is advected by (a) 15 m and by (b) 3 m, respectively. Here, nacc = 50 and nυ = 20. The turbulence parameters are ϵ = 7 × 10−3 m2 s−3, κout = 0.002 m−1, and κin = 0.33 m−1.

  • Fig. 13.

    (a): HDL line of sight (dashed line) from the Laboratoire de Météorologie Dynamique in Palaiseau with respect to the underlying terrain (solid line). (b): SNR as a function of range. The statistical uncertainty on the whole dataset is shown with vertical bars (±1 − σ).

  • Fig. 14.

    Comparison between spectra retrieved from processed actual signals (thin solid line), recorded by 10-μm HDL operated by Laboratoire de Météorologie Dynamique (LMD) during the ECLAP experiment on M13 at 1430 UTC, and predicted by theory (heavy solid line). The turbulence parameters used to compute the theoretical spectrum Φr(κ) are (a) ϵ = 4.6 × 10−3 m2 s−3 and κout = 0.0022 m−1 on the entire range (2.8 km; 7 km), (b) ϵ = 3.7 × 10−3 m2 s−3 and κout = 0.0021 m−1 on (2.8 km; 5 km) range, and (c) ϵ = 5.2 × 10−3 m2 s−3 and κout = 0.0023 m−1 on (5 km; 7km) range. The corresponding turbulence parameters measured by the sonic anemometer and the backscatter lidar on M13 at 1430 UTC are u∗ = 0.67 m s−1, w∗ = 1.49 m s−1, zi = 0.87 km, LMO = −198 m, ϵ = 5.0 × 10−3 m2 s−3, and 3.0 × 10−3 m2 s−3 for Eqs. (37) and (38), respectively.

  • Fig. 15.

    Histogram of outer scale of turbulence κ−1out retrieved by LMD 10-μm HDL during ECLAP experiment. The instrument precision on κout estimates is ≃10%–20%. The natural variability for κout is of the same order of magnitude (see Fig. 14).

  • Fig. 16.

    Time evolution of the TKE dissipation rate ϵ from sonic anemometer measurements at 10 (solid line) and 30 m (dashed line) height on (a) M09, (b) M10, (c) M13, and (d) M14. The TKE dissipation rate is computed from the relationship ϵ = u3[φM(z/LMO) − z/LMO]/kz.

  • Fig. 17.

    Turbulent parameters from sonic anemometer measurements on (a) M09, (b) M10, (c) M13, and (d) M14. The thin solid and dashed lines represent u∗ and w∗, respectively. The thick solid and dashed lines represent ϵ computed from Eq. (37) (Hojstrup 1982) and Eq. (38) (Moeng and Sullivan 1994), respectively.

  • Fig. 18.

    (a) Scatterplot of best-fit spectral model for ϵ [Hojstrup (1992) model with circles and Moeng and Sullivan (1994) model with stars] computed from sonic anemometer measurements vs ϵ retrieved by LMD 10-μm HDL during ECLAP experiment. The instrument statistical uncertainty is shown with vertical bars (±1 − σ). (b) Scatterplot of best-fit spectral spectral model for the velocity standard deviation (Panofsky et al. 1977; Kristensen et al. 1989) computed from sonic anemometer measurements vs σ̂υt computed from Φ̂t(κ) estimated from the inversion procedure. The instrument statistical uncertainty is shown with vertical bars (±1 − σ).

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 255 54 15
PDF Downloads 121 37 0