Vous êtes sur la page 1sur 9

3rd AIAA Atmospheric Space Environments Conference AIAA 2011-3198

27 - 30 June 2011, Honolulu, Hawaii

Characterizing Aircraft Wake Vortices with Ground-Based


Pulsed Coherent Lidar: Effects of Vortex Circulation
Strength and Lidar Signal-to-Noise Ratio on the Spectral
Signature

Daniel J. Ramsey 1 and D. P. Chi Nguyen 2


Aerospace Innovations, LLC, Yorktown, VA 23692

The pulsed Coherent Lidar is a well-known instrument for detecting and characterizing
transverse cross-sections of aircraft wake vortices. One class of algorithms uses spectral
maximum velocities to estimate vortex parameters, and these velocities are dependent
upon several factors described in this paper. A lidar simulator, developed by Lockheed
Martin Coherent Technologies, was used to investigate these factors and explore their
qualitative effects on the spectral signature.
An algorithm is briefly described which attempts to manage these effects, and examples
are shown of vortex tracks resulting from processing simulated and field-measured lidar
data.

Nomenclature
= vortex model core radius
= measurement radius
= circulation strength
= vortex tangential velocity

I. Introduction

A IRCRAFT wake vortex transport and decay is an important topic in air transportation safety given the
increasingly dense air traffic scenarios predicted in the coming decades. Multiple instruments have been used
to study vortex behavior in the field, but the pulsed coherent (Doppler) lidar has proven successful at characterizing
transverse cross-sections of wake vortices at distances ranging from a few hundred meters to a few kilometers.
Several processing algorithms have been developed to extract relevant vortex parameters from lidar data. One
class of algorithms uses the spectral "maximum velocities," 1,6,12 or "velocity envelopes," 7, 10, 13 to extract a single
velocity measurement from each spectrum. From these measurements, the location and strength of the vortices is
then inferred. Another set of algorithms compares an analytical vortex model template to the spectral data, using
some form of matched filter5 or log-likelihood maximization4 in the spectral space.
This paper addresses the former class of algorithms, in particular, the bias inherent in measuring maximum
velocities and the apparent sources of this bias. This investigation of key factors influencing the spectral signature is
ultimately intended to improve the accuracy of real-world wake vortex parameter estimation.
In this work, we used a wind lidar simulator developed by Lockheed Martin Coherent Technologies (LMCT) in
order to explore key factors influencing the spectral signature. Designed to simulate the WindTracerTM 2m pulsed
lidar, the LMCT simulator allows us to vary a number of variables in the vortex wind field and in the sensor system
itself.9 In many cases, we have also used data from NASA field collection campaigns (Denver 2003, John F.
Kennedy 1997) to confirm the same phenomena observed in the simulated data sets.

1
Research Engineer, 4822 George Washington Memorial Highway, Suite 200, AIAA Member.
2
Principal Scientist, 4822 George Washington Memorial Highway, Suite 200, AIAA Member.
1
American Institute of Aeronautics and Astronautics

Copyright 2011 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
II. Maximum Velocity Estimation

A. Spectrum Estimation
The 2m pulsed lidar transmits pulses modulated with a high-
frequency heterodyne oscillation. The backscattered radiation from
aerosols is exhibits a frequency shift due to their velocity
component along the line-of-sight (LOS), which is observed in the
spectrum of the return signal.
Lidar spectra are estimated by sampling the return signal from
each pulse in overlapping range gates. After applying a window
function and computing the discrete Fourier transform for each
gate, the resulting spectra are corrected for frequency jitter by
shifting according to the deviation of the transmitted pulse from the
nominal heterodyne frequency. By accumulating a series of pulses,
the final estimate of the lidar spectra is obtained for each range gate Figure 1. The presumed scanning
along a single LOS. Vertical scanning perpendicular to the aircrafts configuration is a vertical scan plane
flight path results in a planar data collection region (Fig. 1). approximately perpendicular to the
aircraft flight path. The resulting
B. Maximum Velocity Measurement spectral data is mapped to polar
Figure 2 illustrates the components of a lidar spectrum for a coordinates with the lidar as origin.
single sampling volume, whose dominant Doppler velocity is represented by a shift in the spectral peak. The
effective length of the this volume (typically on the order of 100m to 200m) is a result of the combined lengths of
the lidar pulse width and temporal range gate.8 The significant length of this volume relative to vortex dimensions
implies that the peak velocity characteristic is insufficient to adequately characterize vortices. Rather, vortex
tangential velocities are most apparent as a widening of the spectrum.
The maximum velocity approach aims to describe the broadening in
these tails with a single radial velocity value in each direction
(positive and negative).
We define the maximum velocities here as the most extreme
spectral lines above a certain noise threshold. This threshold may be a
fixed value or may be determined locally by a number of methods. In
these examples, the noise is first estimated by measuring the spectral
power amplitude in the outer margins of a bandwidth of interest. The
threshold is then set at a fixed multiple of standard deviations above
this mean noise level.12 The closest intersections of the spectral curve
with the noise threshold define the maximum velocity on each side of
the peak.
It is evident that the threshold-intersection method will always
yield a maximum velocity spread greater than zero, even in the base
case of an air volume undisturbed by aircraft. The width of this system
Figure 2. Schematic of key spectral broadening is due in part to atmospheric effects, scan configuration,
components. The peak velocity represents lidar transceiver characteristics, and inevitable artifacts of discrete
the dominant frequency in the sampling signal sampling and windowing.
volume. Noise levels are measured in the
margins of the spectral bandwidth (here,
14m/s to 24m/s Doppler velocity). A
threshold value is set at a multiple of
standard deviations above the noise
mean, and the intersections of the
spectral curve and the noise threshold are
treated as maximum velocities.

2
American Institute of Aeronautics and Astronautics
Regardless of particular lidar
system characteristics or atmospheric
conditions, these same mechanisms
also induce a broadening in the
spectra that are disturbed by vortices.
This implies that the observed
maximum velocities are reasonably
expected to exceed their theoretical
values by some margin. Left
unmodified, these velocities would
result in a systematic overestimation
of vortex velocities, and thus
circulation.
With access to a realistic lidar
simulator, we can compare these
measured maximum velocities to
those used to generate the simulation.
For a baseline case, we use the
Burnham-Hallock model,2 which
describes the tangential velocities of a
single vortex cross-section at each
measurement radius r as a function of
the total circulation parameter and a
core radius, rc.


= (1)
2 2 + 2

These theoretical velocities are


plotted against the experimentally
derived maximum velocities in Fig. 3
for two circulation values and a range
of SNR.

C. Broadening Regimes and


Circulation Effects
Figure 3. Maximum velocities, calculated from simulated lidar spectra, Pulsed Doppler lidar measures
are compared to tangential velocity profiles of analytic models for a only the LOS (radial) components of
single weak (left) and strong (right) vortex with SNR from 2 dB to 10 atmospheric velocity. Because the
dB (rc = 2.5m). Near-core dropout is clearly present in the strong effective length of the sampling
vortex, demonstrating increasing likelihood with noise. The weaker volume is fixed, the fraction of
vortex does not produce this effect. Increased system broadening with aerosols with velocity vectors
SNR is also visible. In order to clearly demonstrate these effects, an substantially parallel to the LOS
artificially high number of pulses (N=100) was used to reduce random increases with the radius of curvature
variations in these high-density scans. of the vortex streamline. At the same
time, the magnitudes of these
tangential velocities are decreasing. The combined effect of these characteristics on the spectral signature can be
broadly classified in two regimes.
Regime 1 is observed when smaller, faster volumes near the vortex core contribute a small number of high speed
aerosol returns and a large volume of slower returns (along the LOS vector). For a sufficiently strong vortex, the
core velocities will appear in the spectrum as an outlying bump with a peak distinct from the bulk of the aerosol
velocities (Fig. 4, top). If the vortex is not sufficiently strong, however, the slower core velocities will appear as a
broadening of the bulk distribution, usually without a distinct peak.

3
American Institute of Aeronautics and Astronautics
Regime 2 is observed when larger, slower moving masses of air
further from the vortex core exhibit a broadening of the bulk
distribution, without a distinct peak (Fig. 4, bottom). As the
proportion of the vortex mass moving parallel to the LOS increases
with radius, the vortex velocities dominate the sampling volume and
the spectral peak will be shifted from the ambient wind velocity. Due
to their lower velocities, weaker vortices will often exhibit this second
mode of broadening throughout.

III. SNR Effects


The magnitude of the broadening in both regimes is strongly
dependent upon SNR, defined as the ratio of the average heterodyne
signal power to the average noise power.3 As SNR decreases, the
rising noise level (and corresponding threshold) begins to exceed the
magnitude of the broadened tails, eclipsing any information contained
therein and decreasing the measured maximum velocities.
In Regime 1, the characteristic outlying spectral bump has a height
proportional to the size of the corresponding air volume. The effective
length of the sampling volume for pulsed lidar results in a relatively
low-magnitude for these bumps relative to the bulk of the spectrum.
This makes them particularly vulnerable to noise. At the crest of these
bumps, where the slope of the spectral curve is rather flat, small
increases in noise have a large effect on the maximum velocity
estimate.
This highly nonlinear, near-core dropout behavior occurs in
stronger vortices and becomes increasingly prevalent in all vortices in
low-SNR conditions. As is apparent even with the artificially smooth
spectra used to produce Fig. 3, the maximum velocity cannot be
reliably measured inside of a minimum radius (5m to 10m) without
exceptionally high SNR conditions achieved with high pulse power,
longer pulse width, hazy or dusty atmospheric conditions, or smoke-
seeded vortices.
If the precise vortex center is known, however, the velocities in a
range outside of this minimum radius may be used to determine an
average circulation value by averaging the individual circulation

Figure 4. Simulated lidar spectra for a


single Burnham-Hallock vortex at a
range of measurement radius values, r.
Figure 5. Spectra for two range gates near the core of a Boeing (=500m2/s, rc=2.5m, N=100, SNR=10dB).
747 wake vortex. These spectra are from samples above (left) and Measured maximum velocities (vmax),
below (right) the vortex center. Both spectra exhibit outlying indicated by the vertical dotted line,
bump broadening characteristic of high velocities near the consistently exceed the theoretical
vortex core, the first regime described above. The unusually high tangential velocity (vt). Transition from
SNR of this data (15dB to 18dB) permits measurement of these broadening Regime 1 (top) to Regime 2
high-velocity regions. (NASA lidar; 1997 JFK field deployment). (bottom) is plainly visible.
4
American Institute of Aeronautics and Astronautics
estimates for each point (given by 2rvt ,) over a range of radius values. This does not solve the system broadening
bias inherent in Regime 2, nor does it simplify the location of the vortices using maxima in the velocity data.
In Regime 2, the velocity profile is more uniform. Here,
system broadening causes a velocity bias dependent on both
SNR and velocity. The system broadening, measured by
averaging the spectra in the margins of the local scan area, can
be subtracted from the measured velocities to remove SNR-
dependent effects in this regime.1, 6 The velocity dependence
remains, however, and we are left instead with a systematic
underestimation of the velocity values. Fig. 6 illustrates the
absolute error for this method for a single vortex case in
multiple SNR conditions.
Accurately correcting for system broadening in the
maximum velocity space is more complex than simply
removing the ambient velocity spread. This method is a good
Figure 6. Absolute residual error from the start toward normalizing the maximum velocity in Regime 2,
model velocity after subtracting ambient but it does not adequately account for the dependence of this
maximum velocity spread from measured error on measurement radius and tangential velocity.
vortex maximum velocities in a series of
simulation scenarios. Outside of a minimum IV. Vortex Pair Velocity Ambiguity
radius (5m), the velocity error follows a power
law-type behavior; at larger radii, the The examples shown up to this point are reduced to a
spectrum increasingly resembles the ambient single vortex in order to more clearly demonstrate the
case, and the estimation error approaches fundamental spectral characteristics. The more practical task
zero. Near the vortex core, the behavior is of characterizing the full vortex pair is a more complex
more unpredictable from line to line. ( = 500 challenge.
m2/s, rc = 2.5m, N = 100, SNR = 0 dB to 10 dB Accounting for the velocity distortion of a vortex pair,
combined) usually modeled by the superposition of two model vortex
velocity fields, has been thoroughly demonstrated.6,7,10 An
additional inherent problem for pulsed lidar is resolving the velocity contributions from each vortex when the
system range resolution is typically quite low.
The effective length of the range gate sampling volume is a result of the combined lengths of the lidar pulse
width and temporal range gate.8 This total length, typically on the order of 100m to 200m, is significant given the
scale of the vortex pair. Depending on the observation angle, a single volume will typically sample aerosols from
both vortices simultaneously.
Overlapping the range gates by some fraction allows a superficial increase in range resolution at the cost of
increased computational effort. Even so, the inherent range ambiguity due to the volume length remains unresolved.
In angular terms, the scan lines between the vortex cores share the same velocity envelope (high or low,
depending on orientation), and these velocities appear to overlap. As a result, this region is muddled and it is
difficult to discern the individual velocity contributions. Further complicating the problem, the chaotic inner region
is usually comprised of the near-core velocity regions, which are subject to drop-out at low SNR values.
In contrast, the outer tails are distinct and generally maintain the same basic single-vortex characteristics
described in the previous sections. These uncorrupted regions can be used to estimate circulation if the position is
known accurately.
The common overlapped region also makes accurate position estimation more difficult, however. Estimating
position using velocity maxima10, 13 is complicated and ambiguous in this muddled region, and using single vortex
matched filters or other single-vortex templates will result in measurable position and circulation biases due to the
apparent increased velocities in this region.9

5
American Institute of Aeronautics and Astronautics
Figure 7. Illustration of vortex pair velocity ambiguity. These false-color images are made up of the simulated
spectra for every scan angle at a single range gate. Measured maximum velocities are shown here as solid
white lines on either side of the spectral peaks. A large number of pulses averaged (N=100) is used in these
high-resolution scans to minimize random variation. In the center and right vortex pair plots, the apparent
overlap is evident. The outer tails (left side of the spectrum) are distinct and are clearly attributable to each
vortex. The muddled inner region, however, results in position and circulation distortions when using a single
vortex template. In lower SNR conditions (right plot), the near core dropout is more pronounced, which
increases the ambiguity in the inner region.

V. Wake Vortex Parameter Estimation Algorithm


To estimate the basic vortex parameters for simulated and real lidar data, we have developed a processing
algorithm using estimated maximum velocities. This algorithm is loosely based on a NASA algorithm developed
during the AVOSS (Aircraft Vortex Spacing System) studies in the late 1990s and early 2000s 1, 11, 12 with
significant modifications.

A. Algorithm Description
A movable region of interest (ROI)14 consists of a restricted set of scan lines and range gates within which the
algorithms effort is contained, and the tracking algorithm continually attempts to center the ROI on the vortex
position. This region is used to differentiate between vortex pairs from multiple aircraft, as well as to define the
local ambient conditions, including SNR, noise threshold, and system broadening. The ROI initial position is the
location where the vortex pair is likely to appear (the flight corridor directly over the runway for landing or
departing aircraft), and follows the vortices movement around the greater scan frame. To determine the ambient
conditions, samples are taken from range gates at the margins of the scan region, where the vortex does not
influence these properties.
The maximum velocities are estimated as described in Section II, which results in a pair of maximum velocity
maps for the ROI (both high and low velocity envelopes). The system broadening is removed by subtracting the
ambient velocity spread from each value in the ROI.
A convolution is then performed on these maps simultaneously, using an analytical vortex template with points
in two spatial dimensions as well as in the high and low velocity envelopes. The values of the template are chosen
such that the resulting product is proportional to the model circulation at each point in the ROI. The highest
response for the positive (near) and negative (far) vortices are chosen as the initial locations, and the magnitude of
their peaks is chosen as the initial circulation estimate.
With the knowledge that the initial position estimate is influenced by the presence of the opposing vortex
(Section IV), the second iteration improves the position estimate by searching for local maxima in the outer tails of
the velocity envelopes for each vortex.
A multiple-target tracking algorithm initializes, maintains, and terminates vortex tracks using field data files that
may contain multiple aircraft flybys. The end result is a series of track files in a standardized format containing

6
American Institute of Aeronautics and Astronautics
circulation and position estimates (lateral and vertical) for each scan in which the track is detected. Additionally, a
number of visualization products enable a close examination of the data during and after processing.

B. Selected Track Results


This algorithm, currently still in development, has been used to process a number of simulated data sets and a
limited set of field data. Selected track data from these studies are shown in Fig. 8, 9, and 10.

Figure 8. Track results from simulated lidar data using wind a wind field generated by a NASA large eddy
simulation (LES). The LES is intended to simulate a Boeing 747 wake vortex pair in weak turbulence and no
crosswind. The dashed lines correspond to the simulation truth data. (SNR = 10 dB.)

7
American Institute of Aeronautics and Astronautics
Figure 9. Track results from a Boeing 777. WindTracer spectra from the 2003 Denver study were processed
using the algorithm described above to produce these results. (SNR = 5 dB.)

Figure 10. Ensemble track results from a 30-minute period during the 2003 Denver study, which includes
wakes from a diverse aircraft mix. (SNR 4 dB to 6 dB).

8
American Institute of Aeronautics and Astronautics
VI. Conclusion
The effects discussed here are strong influences on the vortex signature, particularly when using maximum
velocity techniques to characterize vortices. The location, size, and velocity distribution of the measured air volume,
as well as SNR, all affect the error of the maximum velocity measurement with respect to the true analytical
velocity value. Failure to account for these effects will result in systematic inaccuracies when measuring vortices in
the field. The algorithm introduced above attempts to manage these effects and has been successfully demonstrated
on simulated and real-world field data. Preliminary testing indicates that this algorithm is most effective in high-
SNR conditions and with higher scan densities, but ongoing research is underway to increase robustness in the
lower-density and low-SNR data sets produced by currently fielded sensor systems.

Acknowledgments
This work was sponsored by the National Aeronautics and Space Administration Air Space Systems Program.
The work was completed under the NASA NRA Enabling Super-Dense Operations by Advancing the State of the
Art of Fast-Time Wake Vortex Modeling. This research was funded through Northwest Research Associates under
contract NNL08AA45C. The authors would like to thank Neil OConnor of NASA Langley Research Center as
well as Don Delisi and David Lai of NWRA for their advice and support. The authors also wish to thank Don Jacob
of Lockheed Martin Coherent Technologies, whose lidar simulator was used extensively in this research.

References
1
Britt, C. L., Estimation of Aircraft Wake Vortex Size and Strength from Coherent Pulsed Lidar Measurements,
Proceedings of the 9th Conference on Coherent Laser Radar, 1997, pp. 122-124.
2
Burnham, D.C., and Hallock, J.N., "Chicago Monostatic Acoustic Vortex Sensing System," Report No. DOT-TSC-FAA-
79-103. IV, July 1982, 206 pp.
3
Frehlich, R., Effects of Wind Turbulence on Coherent Doppler Lidar Performance, Journal of Atmospheric and Oceanic
Technology, Vol. 14, 1997, pp. 54-75.
4
Frehlich, R. and Sharman, R., Maximum Likelihood Estimates of Vortex Parameters from Simulated Coherent Doppler
Lidar Data, Journal of Atmospheric and Oceanic Technology, Vol. 22, 2005, pp. 117-130.
5
Hannon, S. M. and Thompson, J. A., Aircraft Wake Vortex Detection and Measurement with Pulsed Solid-State Coherent
Laser Radar, Journal of Modern Optics, Vol. 41, No. 11, 1994, pp. 2175-2196.
6
Heinrichs, R. M. and Dasey, T.J., Analysis of Circulation Data from a Wake Vortex Lidar, AIAA Meeting Papers on Disc,
January 1997, A9715151, AIAA Paper 97-0059.
7
Holzpfel, F., Gerz, T., Kpp, F., Stumpf, E., Harris, M., Young, R., and Dolfi-Bouteyre, A, Strategies for Circulation
Evaluation of Aircraft Wake Vortices Measured by Lidar, Journal of Atmospheric and Oceanic Technology, Vol. 20, 2003, pp.
11831195.
8
Huffaker, R. M. and Hardesty, R. M., Remote Sensing of Atmospheric Wind Velocities Using Solid-State and CO2
Coherent Laser Systems, Proceedings of the I12, Vol. 84, No. 2, Feb. 1996.
9
Lai, D. Y., Jacob, D., and Delisi, D. P., "Assessment of Pulsed Lidar Measurements of Aircraft Wake Vortex Position
Using a Lidar Simulator," Paper AIAA 2010-7988, AIAA Atmospheric and Space Environments Conference, Toronto, Ontario
Canada, Aug., 2010.
10
Kpp, F., Rahm, S., and Smalikho, I., Characterization of Aircraft Wake Vortices by 2-m Pulsed Doppler Lidar,
Journal of Atmospheric and Oceanic Technology, Vol. 21, 2004, pp. 194206.
11
Nguyen, D.P.C., Wake Vortex JFK2 Final Report: May 26 - June 6, 1997 JFK2 Deployment: Wake Vortex Measurements
at John F. Kennedy International Airport for a 2.02 mm Pulsed Coherent Lidar, a 10.6 mm Continuous Wave Lidar, and a
Ground Wind Vortex Sensing System, Research Triangle Institute, Center for Aerospace Technology, 1998.
12
Nguyen, D. P. C. and Britt, C. L., Wake Vortex Lidar, NASA Lidar 1997 Dallas/Fort Worth Field Measurements and Data
Guide, RTI/4903-032-04, Research Triangle Institute, Hampton, VA, May 1998.
13
Rahm, S., and Smalikho, I., Aircraft Wake Vortex Measurement with Airborne Coherent Doppler Lidar, Journal of
Aircraft, Vol. 45, No. 4, 2008, pp.1148-1155.
14
Switzer, G. F., Proctor, F.H., Ahmad, N. N., Duparcmeur, F.M.L., An Improved Wake Vortex Tracking Algorithm for
Multiple Aircraft, Paper AIAA 2010-7993, AIAA Atmospheric and Space Environments Conference, Toronto, Ontario Canada,
Aug., 2010.

9
American Institute of Aeronautics and Astronautics

Vous aimerez peut-être aussi