1. Introduction
After roll-up of the wake of an aircraft, a pair of counterrotating vortices is created. This phenomenon, called wake vortex (WV) turbulence, can constitute a hazard for any other aircraft behind the first, especially during takeoff and landing phases (Krause 2003). Indeed, the aircraft encountering a wake vortex system that has not dissipated enough may be subjected to a strong rolling moment (when encountering one of the vortices) or to a strong downwash velocity (when flying between the two vortices), causing a drop in lift.
To prevent such risks of encountering a wake, safety separations between aircraft were imposed by the International Civil Aviation Organization (ICAO) in the early 1970s. With the current exponentially growing air traffic, those safety separations have become too constraining for large and busy airports, limiting their capacity and flexibility, and causing many delays in departures and arrivals.
To optimize the airport/runway throughput while maintaining an acceptable level of safety, the separation minima were redefined in 2013 by splitting the ICAO heavy and medium categories into “upper” and “lower” aircraft parts, resulting in six categories instead of the four previous ones (Rooseleer and Treve 2013). This new categorization [European Wake Vortex Recategorisation (RECAT-EU)] has been deployed in France at Paris Charles de Gaulle airport since the spring of 2016 (Direction Générale de l’Aviation Civile 2016). A similar system (RECAT) has been implemented by the Federal Aviation Administration (FAA) in the United States (Cheng et al. 2016).
To support the development of new systems for safe decrease of aircraft spacing (such as weather-dependent separation systems), measurement tools that can provide information about the wake vortex situation around the airport in real time will also be required. The main sensors used for this issue are lidar and radar. Lidars operate in clear air but they require a sufficient concentration of aerosols in the air and they are blind in rain or fog conditions. Radars are able to see scattered electromagnetic waves on raindrops or fog particles but the power of the scattered electromagnetic waves is very weak in clear air around 10 GHz. To cover all weather conditions, both tools will likely be used together (Oude Nijhuis et al. 2018). For reliable detection, it is necessary to know what the limits of detectability for the sensors are under different weather conditions. The model presented in this paper proposes a new tool for the optimization of the radar parameters to detect WV in clear-air conditions. The analysis is based on using 3D fields from large-eddy simulations (LES) for generating the time series and the Doppler spectra of the scattered signal.
Detection of a wake vortex under clear-air conditions was demonstrated for VHF, UHF, L, S, and C bands (Noonkester and Richter 1980; Chadwick et al. 1984; Nespor et al. 1992; Gilson 1992; Shephard et al. 1994) and for X band (Shephard et al. 1994; Barbaresco et al. 2007, 2008), with an unsuccessful attempt reported in Marshall et al. (1997). The weak level of the backscattered power is the origin of the difficulties in detecting the WV with the radar at large distances. Radars provide remote sensing of the atmospheric phenomena at very large ranges in C and S bands. However, the long range is not the main requirement for the system aimed at the detection of the WV. Among the main advantages provided by the X band are high resolution relative to the C and S bands and a smaller sensor size. Therefore, the use of a wide bandwidth in the X band provides high-resolution measurements. Additionally, because of the smaller wavelength, it is possible to obtain a narrow beamwidth using a relatively small antenna.
Models describing the scattering mechanisms of the electromagnetic waves on the wake vortex are reported in the literature. They can be classified into two groups: the first one assumes the scattering on the spiral density pattern of the WV evolving with time (Myers et al. 1999; Shariff and Wray 2002; Li et al. 2010; Vanhoenacker-Janvier et al. 2012) and the second one assumes the scattering on the turbulence induced by the WV (Marshall and Myers 1996). The main limitations are that the models available are 2D models and most of them do not take into account the turbulent nature of the interaction between the WV and the surrounding turbulent atmosphere.
Vanhoenacker-Janvier et al. (2012) have developed an electromagnetic model for a radar cross section (RCS) due to wake vortices, as a function of meteorological and airplane parameters, using the Li et al. (2009, 2011) oscillatory integral method. This first model used a simplified 2D model to simulate the wake vortices. In this study, the electromagnetic simulator uses the stratified and turbulent wake vortices output from a state-of-the-art LES, thus providing more realistic results. An improvement of the computing time for calculation of the backscattered signal time series is obtained, by using the Muschinski et al. (1999) model. The four main assumptions of the Muschinski et al. (1999) model are that 1) the size of the radar resolution volume is much larger than the LES grid resolution, 2) the size of the LES cells is much larger than the Bragg wavelength, 3) the turbulence is locally isotropic and homogeneous at the scale of the LES cell, and 4) the LES subfilter turbulence belongs to the inertial subrange of the refractive index spectrum on the Bragg scale. The first three assumptions are valid for the parameters of the modeled X-band radar and the LES resolution used in this study. However, with regard to the fourth assumption, it is expected that the Bragg scale lies in the dissipative range or in the transition zone from the inertial subrange to the viscous dissipative range for the X-band radar, which necessitates a more accurate modeling of the refractive index spectrum. To take this transition into account, the Hill (1978a) spectral model is applied in this study.
Section 2 of this paper describes the large-eddy simulation of the WV, the input parameters chosen, and illustrates the time evolution of various WV parameters and atmospheric fields. Section 3 describes the model developed for the calculation of the radar backscattering in clear air and further describes the hypotheses used for the application at X band. Section 4 analyzes the impact of the relative size of the smallest turbulent eddies with respect to the Bragg wavelength. Section 5 develops the method used for the calculation of the Doppler spectrum of the signal received by the radar, and section 6 demonstrates the estimated WV circulation using two methods: from the Doppler spectrum and from the estimated WV descent velocity and the distance between WV cores.
2. Large-eddy simulation of the wake vortices
The aircraft wake vortex system obtained after roll-up, as simulated here, consists of a pair of counterrotating vortices with circulation Γ0 and −Γ0 and separated by a distance b0. The aim is to simulate those vortices in a stratified and turbulent atmosphere, and with the presence of a humidity gradient. This setup is similar to that of De Visscher et al. (2013), with the addition of the humidity field. The simulations were carried out using a pseudospectral solver. A homogeneous isotropic turbulence is first generated and converged to create a realistic background stratified turbulence. Next, the vortex pair generated by a heavy aircraft is superimposed over the turbulence.
a. Governing equations
The computational box size is set to 4b0, which corresponds to a few hundred meters for a heavy aircraft (i.e., b0 on the order of 50 m). Regardless of the vortex wake simulated, the computational domain is too small to consider the Coriolis term (Rossby number ≪ 1) in the momentum equation; therefore, it has been neglected.
b. Numerical method
To solve those equations, the same pseudospectral Navier–Stokes solver as in De Visscher et al. (2013) is used here with phase shift partial dealiasing (Canuto et al. 2006; Jeanmart 2002). The governing equations are thus solved in a periodic domain. The main advantage of spectral solvers is that they are free of dispersion and diffusion errors. The time integration is carried out using Williamson’s (1980) third-order, low-storage Runge–Kutta scheme. The SGS model used for the simulations is the regularized variational multiscale model (Jeanmart and Winckelmans 2007): it dissipates only at the smallest resolved scales of the LES, thus allowing the dynamics of the large and medium scales (the most important ones) to be preserved. The temperature and humidity SGS fluxes are obtained assuming unity SGS Prandtl and Schmidt numbers as in (Sagaut 2006).
c. Numerical setup
In the present simulations, atmospheric conditions encountered during approach are considered, that is, a very weakly/weakly stratified atmosphere combined with atmospheric turbulence ranging from weak to strong. Note that very strong unstable conditions may also be encountered in the lower layers of the atmosphere, yet the periodic setup used here does not allow us to simulate such unstable conditions.
To simulate the wake vortices in realistic conditions, background turbulent fields for each case have been generated first and converged statistically in the computational domain. Those fields were converged using the Alvelius (1999) forcing term, which is characterized by a random phase and an amplitude computed according to a prescribed injection spectrum (Vanhoenacker-Janvier et al. 2014). The integral of this spectrum corresponds to the amount of energy injected per unit time dEf/dt = εf, and thus also to the dissipation rate when the stratified turbulence is statistically converged.
Once the stratified turbulence has been produced, the two counterrotating vortices (aligned with the x direction) are superimposed over the background turbulent velocity field. A low-order algebraic profile [also called Burnham–Hallock (B-H) in the aircraft community] is used as the initial profile of the vortices. The effective core size rc (i.e., the radius of maximum induced velocity) is set to rc/b0 = 0.05, which is typical of what is used when modeling aircraft wake vortices with that model. The Reynolds number based on the vortex circulation is very high: ReΓ = Γ0/ν ≃ 107.
The box size L is 4b0, with a uniform grid size Δ of resolution equal to b0/Δ = 64; rc/b0 = 0.05 has been chosen for the core size ratio, because such a resolution ensures that the most relevant dynamics of the vortices are well captured (rc/Δ > 3) (De Visscher 2012). For each case, the simulation was run for two additional characteristic times t0, after reaching the statistical turbulent equilibrium of the vortices with the background turbulence [=time of maximum eddy dissipation rate (EDR); Fig. 1].
The three cases used in the radar simulator are summarized in Table 1. The dimensional/physical parameters were here obtained by assuming a heavy aircraft (Γ0 = 400 m2 s−1; b0 = 50 m), an initial height z0 of 500 m, and typical atmospheric conditions; thus, V0 = 1.27 m s−1 and t0 = 39.3 s.
Simulation parameters of the three simulated cases used for the radar simulator. The subscripts g and 0 denote atmospheric factors at ground level (z = 0) and at the release height z0 of the two vortices, respectively.
d. Results
Once the vortices are introduced in the simulation, the vortex pair starts to descend at a velocity V0 = Γ0/(2πb0). The turbulence surrounding the vortices is affected by two main phenomena. First, the background turbulence is stretched by the vortex system, generating intense transverse vortical structures. Second, due to the stratification, baroclinic vortical structures are generated at the border of the Rankine oval. This extra source term of vorticity arises from the difference in density between the warmer fluid carried by the vortices inside the Rankine oval and the cooler fluid outside. Subsequently, small-scale instabilities develop inside those structures, leading to a fully turbulent vortex system. The turbulence and the associated EDR increase until reaching a statistical equilibrium with the vortex system (Fig. 1). From that moment on, the flow becomes representative of a real vortex system as obtained after roll-up. This physical evolution is highlighted in red on the graphs. After reaching this equilibrium, the circulation of both vortices decreases smoothly (Fig. 2a), with the vortex system sustaining its own turbulence. Therefore, the descent velocity as well as the maximum tangential velocity (Fig. 2b) of the vortices also decreases. This equilibrium is reached at time t* = t/t0 = 2.5 for cases 1 and 3 and t* = 4 for case 2.
In the LES, the total circulation of each vortex
We see that the total circulation has already decayed greatly, as shown in Fig. 2a, which depicts the time evolution of the dimensionless vortex circulation
The stratification also creates baroclinic vortical structures. Those structures quickly interact with the turbulent vortices, enhancing the decay of the vortices compared to the less stratified case at same level of turbulence, that is, case 1 (Fig. 2a). This baroclinic vorticity also has an opposite sign relative to the primary vortices and thus tends to slow down the vortices as seen in Fig. 2c, which depicts the dimensionless vertical position of the vortex tubes:
For nonstratified or weakly stratified atmosphere, the circulation is proportional to the product of the descent velocity Vc and the distance between the WV cores bc. Figure 2d illustrates this ratio: for case 1 the circulation calculated from the WV velocity field is equal to that calculated from Vc and bc, whereas for the more stratified cases 2 and 3 the ratio deviates significantly from unity.
Figure 5b shows the humidity field combined with the wind velocity (w, u) of case 1 at time t* = 3.75 in a vertical cross section of the computational box. The black crosses represent the center of the vortex tubes. These were located by identifying the location of the minimum pressure. During its descent, the vortex system carries the lower humidity, which is situated at a high altitude, downward. Such a pattern in the humidity field (as well as in the temperature and pressure fields) may backscatter a characteristic return signal to the radar, which could be used to detect the vortex pair and possibly estimate its intensity as well.
3. Backscattering of the radar signal on wake vortices in clear air
The output of the LES of wake vortices described in the previous section is used for the simulation of the radar signal and its Doppler spectrum. The pulse Doppler radar used during the Seventh Framework Program (FP7) Ultrafast Wind Sensors (UFO; http://www.ufo-wind-sensors.eu/home) project was selected for use in this paper. However, the simulator can be used with other types of radars, in which case, only the scattered signal time series need to be modified in the calculation. Preliminary validations of the simulator were performed by extracting the EDR of the atmospheric turbulence under clear-air conditions with a radar wind profiler (Kovalev et al. 2016) and a scanning radar (Kovalev and Vanhoenacker-Janvier 2017) used during the UFO measurement campaign. This section presents an improved version of the simulator, which was also used for the estimation of the circulation of the wake vortices.
a. Geometry and parameters of the radar used for the simulation
The main parameters of the radar are listed in Table 2. Figure 3 illustrates the structure of the radar signal: during the dwell time, the radar emits 512 pulses with a pulse repetition time of 50 μs. The radar receives the pulses scattered by a given radar cell located at a distance of cτ/2, where τ is the time necessary for an electromagnetic wave to travel from the radar to the scatterer and back and c is the speed of light. A pulse received at time mT + τ, where m is the consecutive number of the pulses in the burst, corresponds to the pulse emitted at time mT and scattered back onto the refractive index inhomogeneities. For the sake of simplicity, the pulses in Fig. 3 are given in the form of a voltage that modulates the carrier frequency of the radar; the carrier frequency is not represented. It is assumed that the modulation of the pulses provides the needed range resolution.
Parameters of the simulated X-band radar.
Figure 4 illustrates the geometry used for the simulation: the radar is located at the side of the glide with the distance r to the runway axis equal to 1100 m. Two geometrical setups are used in the simulator. For setup A, the size of the radar resolution volume in range (Δrrc) is 60 m and the elevation angle of the antenna beam is fixed at 5.3° to monitor the WV at the altitude of the glide (h1 = 100 m). For setup B, the range resolution is 6 m and the radar performs a scan of the LES box in elevation (with a step of 0.27°) to track the WV descent.
b. Calculation of the radar signal
The discrete nature of the output of the LES imposes the transformation of the integral into a sum over the radar cell volume for which the product
Because the LES output data are 3D snapshots of meteorological parameters at discrete time values, time t in Eq. (16) is replaced by the discrete values of kT, where T is the pulse repetition time of the radar.
Taylor’s hypothesis of the “frozen turbulence” (Taylor 1938) is used: because of the relatively short dwell time (25.6 ms), it is assumed that the scattering pattern within the LES cell remains constant and is advected in time by the local velocity field;
The Muschinski et al. (1999) model uses two main hypotheses:
First, the LES subgrid turbulence obeys the Kolmogorov scaling; that is, it belongs to the inertial subrange of the turbulence spectrum, at the Bragg scale.
Second, the initial phases φ0p of the LES grid cell contributing to the sum over the radar cell are statistically independent.
c. Calculation of n and
The
An example of the distribution of
d. Calculation of the phase
Estimation of the phase correlation time scale τϕ, the inner scale l0, and the broadening of the Doppler spectrum Δσυ from the subgrid turbulence contribution.
As discussed before, the developed and turbulent WV have much higher EDR values than those used as the forcing background turbulence. For all cases, the mean value of the EDR increases to 3.9 × 10−3–8 × 10−3 m2 s−3, which corresponds to τφ changing in the range from 35 to 27 ms. Proper use of Eq. (19) necessitates a random change of the phases of Bragg scatterers after time intervals of τφ.
e. Estimation of the subgrid turbulence contribution to the Doppler spectrum width
Table 3 shows the broadening of the Doppler spectrum Δσυ caused by the subgrid turbulence. These estimates are given for the background forcing turbulence as well as for the strong increase of the EDR caused by the developed turbulent WV, which results in an increase of Δσυ for all cases.
4. Limits of detectability
For higher radar frequencies, there are no significant eddies with dimension on the order of the Bragg wavelength. The inner scales of the turbulence injected in the background stratified atmosphere for the LES used in this study are listed in Table 3. For the X-band radar used, the Bragg wavelength belongs to the transition region between the inertial subrange and the dissipative range. Therefore, the hypothesis used by Muschinski et al. (1999) to deduce Eq. (16) is not fulfilled for all the three cases considered. However, the developed turbulent WV system has local EDR values in the WV oval that are much larger than that of the initial forcing turbulence. It is also the case in the background part, outside of the WV oval and downstream. An example of the EDR values in a cross section of the LES box is provided for case 1 in Fig. 6. For an initial turbulence forcing with
As seen from Fig. 7, for the values of ε typical for the WV oval, the Bragg wavelength of the X-band radar is in proximity to the bump value of the attenuation factor. For the lowest EDR, corresponding to the background forcing turbulence of case 2, the attenuation of the refractive index variance spectrum estimated by Hill’s and Tatarskii’s models are −17.5 and −11.5 dB, respectively. In this study, the contribution of each LES cell in Eq. (16) is weighted by the local factor [g(ηkB)]1/2 to take into consideration the shape of the refractive index variance spectrum at the transition between the dissipative and the inertial range. In Fig. 8, an example of the factor g(ηkB) is plotted for the snapshot in case 1 (Fig. 6). This figure clearly shows an increase of the refractive index variance spectrum in the WV oval relative to the background atmosphere.
5. Calculation of the Doppler spectrum
Two geometrical setups, as described hereinafter, are used to simulate the WV detection and its parameters estimation. First, the influence of the noise is not considered in the simulations so as to estimate the possibility of the WV circulation retrieval from the radar measurements. However, low RCS of the WV turbulence limits the detection of the WV on long distances. To evaluate this limitation, the signal-to-noise ratio (SNR) is estimated for the radar parameters used in this study (Table 2) in section 6c.
a. Geometrical setup A
Figure 9 shows the details of geometrical setup A: the radar cell is aligned with the WV core center; the size of the radar cell in range is approximately 60 m; the half-power width of the antenna beam is 3.1°. A decision was made to position the WV core at the center of the radar cell here so as to be able to compare the WV circulation estimated from the Doppler spectrum to the exact value as measured in LES (section 6). At each LES time step, the LES box is placed at the same altitude as defined in section 3a. This compensation for the WV descent provides equivalent conditions for the estimation of the WV circulation.
Figure 10 represents a distribution of the local velocity (intensity) and
b. Geometrical setup B
Geometrical setup B is similar to setup A but the altitude of the LES box tracks the altitude of the descending WV. The radar performs a scan of the LES box in elevation with a step of 0.27°. It is assumed that the position of the LES box does not change during each scan. The range resolution is 6 m. For each radar resolution volume, the Doppler spectrum is estimated with Eq. (28).
Figure 12 illustrates the distribution of the Doppler spectra width among radar cells in one elevation scan. The position of the WV cores is estimated by the local maxima of the Doppler spectrum width (Fig. 13). There are two causes of the discontinuities in estimated coordinates (y, z): 1) a relatively large radar range resolution (6 m) and 2) the fact that the radar cell with the maximum Doppler spectrum width does not necessarily correspond to the actual location of the WV core. To reduce the variance of the estimated WV coordinates, the width of the Doppler spectra in a given radar cell is obtained as an average over 25 closest radar cells in elevation and range (in an area of 5 × 5 of the radar resolution volumes centered at a given radar cell).
6. Estimation of the wake vortex circulation
Two methods are used to estimate the WV circulation. For geometrical setup A, the circulation of the WV is estimated from the distribution of the radial velocities within the radar resolution volume. Those distributions are represented by the Doppler spectrum of the backscattered signal. For geometrical setup B, the estimation is based on the tracking of the WV descent velocity and distance between WV cores.
a. Setup A
However, the Rubin model assumes that the wake vortex circulation Γtot is linked to Vmax by Γtot = 2πrcVmax and thus a top-hat wake vortex model (i.e., uniform vorticity within rc and zero outside). In the B-H vortex model used as the initial condition in this study, Γ(r) varies smoothly from 0 to Γtot and is equal to Γtot/2 at rc. Although the circulation distribution of the developed turbulent WV no longer follows the simple B-H model, it is found that Γ(rc) is also roughly equal to Γtot/2, which holds fairly true for the representative time intervals; see the red parts of Fig. 14: Γ(rc) deviates from Γtot/2 by approximately ±12% for all cases. To take this variation into account, it is proposed to modify Eq. (30) by using Γtot/2 ≃ 2πrcVmax. The coefficient k was evaluated using rc = 0.05b0 = 2.5 m. The actual change of rc over time with respect to the initial spacing between WV cores b0 is shown in Fig. 15: rc deviates from the initial value by from +10% to −20% depending on the LES case. As compared with Rubin (2000), the interval of integration in Eq. (30) is changed to [−Vmax, −Vmin] and [Vmin, Vmax] because the Doppler spectrum is not symmetric: the negative and positive parts of the spectrum represent different parts of the wake vortex system.
To compare the circulation retrieved from a given radar cell to the circulation measured from the LES velocity field, the radar cell center is aligned with the vortex core center. The LES reference Γ(xi, rmax, t) is sought at each xi, where rmax is equal to one-half of the radar cell width in range. Then it is averaged over xi (within the radar resolution volume) to obtain Γtot(t). The total circulation shown in Fig. 16 is the average of the left and right wake vortex circulations, normalized by Γ0. The wake vortices decay in time, dissipating their energy by interacting with each other and with the surrounding turbulent background. Consequently, the WV circulation decreases in time as shown in Fig. 16 for the considered LES cases.
The radar estimates of the circulation at each LES time step are obtained from the radar cells centered at the initial positions of the left and right WV cores (at t* = 0). Then the average between the left and right WV estimates is calculated. The results provided in Fig. 16 show a good agreement with the circulation calculated directly from the LES field. Note that in the LES t* = 0 corresponds to the time of the injection of the wake vortices into the initial stratified turbulence. The moment at which the LES reaches a statistical equilibrium—the state at which the energy dissipation rate reaches its maximum—indicates the time after which the LES provides a physically meaningful simulation of fully developed turbulent wake vortices in the atmosphere. Those moments differ among the LES cases, and consequently, the estimation of Γtot starts accordingly (the red parts in the previous figures, lasting for a total time of 2t0).
The limits of integration (Vmin) in Eq. (30) significantly influence the accuracy of the estimation of Γtot. The lower limits Vmin are chosen so as to exclude the influence of the Doppler spectrum broadening caused by the turbulence induced by the WV: 1.15 m s−1 for cases 1 and 3 and 1 m s−1 for case 2. The values for Vmin are estimated by minimizing the error between the estimate Γradar and the reference circulation from the LES Γtot. However, in real systems, Vmin should be estimated from the radar signal backscattered by the turbulence. For all cases, a constant upper limit Vmax = 12.8 m s−1 is used. Nevertheless, the value of Vmax at each time step can be accurately estimated from a total width of the Doppler spectrum (Fig. 17). The evolution of the Doppler spectrum is shown in Fig. 17 for case 1: the total width of the Doppler spectrum decreases with time; the width of the central part represents the velocity variance related to the turbulence.
To evaluate the influence of the misalignment of the radar resolution volume center relative to the WV core, the location of the radar cell center is changed in range from −45 to 90 m. For each shift in range, the circulation within a given radar cell is computed and Eq. (31) is used to calculate the error of the estimation (the right panel of Fig. 18). It can be seen that this error is minimal as long as the shift of the radar cell center is within ±15 m relative to the center of the WV core. Deviations in the local minima of the NRMSE from the initial positions of the WV cores (range shifts of 0 and 50 m for the left and right WV, respectively) are due to the fact that the wake vortices do not remain symmetrical and centered at their initial positions as they evolve in the turbulent stratified atmosphere.
b. Setup B
We note that De Visscher et al. (2013) showed that the influence of the stratification can be accurately modeled. For cases with significant stratification, the model they proposed could potentially be used for estimating the WV circulation from the measurement of the WV descent velocity.
For geometrical setup B, the coordinates of the WV cores are estimated as explained in section 5b. Then the distance between the WV cores is computed and the descent velocity is estimated using three successive measurements for each WV core. Figure 19 shows the results of the circulation estimation obtained using Eq. (32). The parameters
c. Impact of the noise on the WV detection
Low RCS of the WV limits the radar detection on long distances. Figure 21 shows the SNR estimated for one pulse for different combinations of emitted peak power Ptx and range r. The value of −85 dB is taken for the WV RCS and the noise level of the receiver is −140 dB. The other parameters of the radar are given in Table 2. For the distance of 1050 m and the emitted peak power of 600 W, the single-pulse SNR is approximately −18 dB. However, it is estimated that 7–10 dB of the single-pulse SNR is needed for the accurate retrieval of the WV circulation from the Doppler spectrum (geometrical configuration A). For this SNR, the side parts of the Doppler spectrum (Fig. 11) are above the noise floor. So, for the emitted power of 600 W, the WV circulation can be retrieved at the distance of approximately 250 m. To increase the range, the emitted power should be increased correspondingly.
7. Conclusions
This paper presents a new application of the Muschinski et al. (1999) method and Hill’s (1978b) spectral model to LES output data of wake vortices. The LES data serve as input for the simulation of the backscattered radar signal, based on the Bragg scattering by the turbulent field created by the WV. The ratio between the Bragg wavelength of the radar and the inner scale (itself proportional to the Kolmogorov scale) is an important parameter for the detectability of WV in clear air. The simulation of the radar signal backscattered by the WV, and the associated turbulence, is used for the calculation of the Doppler spectrum, and two models for the retrieval of the circulation are tested. The Doppler spectrum components containing the useful information on the WV are located at approximately 20 dB below the peak of the spectrum. Therefore, the circulation can be estimated fairly accurately if the level of backscattered signal is high enough above the detection threshold of the radar. Moreover, for cases with low stratification, the circulation can also be estimated well using the measured descent velocity and spacing between the vortices. The developed platform and the tested models/methods, constitute a useful tool for the optimization of the radar configuration and parameters and for the improvement of the link budget. It could also be used, alongside other algorithms, for a benchmarking of methods for the estimation of WV circulations in clear air.
Acknowledgments
This research project has been partially funded by the UFO (European FP7 Collaborative Project Grant Agreement 314237), SESAR JU (Project 12.2.2) Thales subcontract agreements. Author Kovalev was supported by an FSR Research Grant from Université catholique de Louvain. We thank two anonymous reviewers for their comments, which helped to improve the paper.
REFERENCES
Alvelius, K., 1999: Random forcing of three-dimensional homogeneous turbulence. Phys. Fluids, 11, 1880–1889, https://doi.org/10.1063/1.870050.
Barbaresco, F., A. Jeantet, and U. Meier, 2007: Wake vortex detection and monitoring by X-band Doppler radar: Paris Orly radar campaign results. Int. Conf. on Radar Systems, Edinburgh, United Kingdom, Institution of Engineering and Technology, https://doi.org/10.1049/cp:20070559.
Barbaresco, F., J. P. Wasselin, A. Jeantet, and U. Meier, 2008: Wake vortex profiling by Doppler X-band radar: Orly trials at initial take-off and ILS interception critical areas. Radar Conf., Rome, Italy, Institute of Electrical and Electronics Engineers, https://doi.org/10.1109/RADAR.2008.4721113.
Bean, B. R., and E. J. Dutton, 1966: Radio Meteorology. National Bureau of Standards Monogr., No. 92, U.S. Government Printing Office, 435 pp.
Blackman, R. B., and J. W. Tukey, 1958: The measurement of power spectra from the point of view of communications engineering—Part I. Bell Syst. Tech. J., 37, 185–282, https://doi.org/10.1002/j.1538-7305.1958.tb03874.x.
Canuto, C., M. Y. Hussaini, A. Quarteroni, and T. A. Zang, 2006: Spectral Methods: Fundamentals in Single Domains. Scientific Computation Series, Vol. 22, Springer-Verlag, 581 pp., https://doi.org/10.1007/978-3-540-30726-6.
Chadwick, R. B., J. Joradan, and T. Detman, 1984: Radar cross section measurements of wingtip vortices. Proc. Int. Geoscience and Remote Sensing Symp., Strasbourg, France, European Space Agency, 479–483.
Cheng, J., J. Tittsworth, W. Gallo, and A. Awwad, 2016: The development of wake turbulence recategorization in the United States. Eighth Atmospheric and Space Environments Conf., Washington, DC, American Institute of Aeronautics and Astronautics, https://doi.org/10.2514/6.2016-3434.
Crow, S. C., 1970: Stability theory for a pair of trailing vortices. AIAA J., 8, 2172–2179, https://doi.org/10.2514/3.6083.
De Visscher, I., 2012: Interaction of wake vortices with the atmosphere and the ground: Advanced numerical simulation and operational modeling. Doctoral thesis, Ecole Polytechnique de Louvain, 339 pp., http://hdl.handle.net/2078.1/110734.
De Visscher, I., L. Bricteux, and G. Winckelmans, 2013: Aircraft vortices in stably stratified and weakly turbulent atmospheres: Simulation and modeling. AIAA J., 51, 551–556, https://doi.org/10.2514/1.J051742.
Direction Générale de l’Aviation Civile, 2016: Implementation of the RECAT-EU wake turbulence separation scheme at Paris-Charles de Gaulle, Paris-Le Bourget and Pontoise–Cormeilles-en-Vexin airports from March 22nd 2016. Direction des Services de la Navigation Aérienne Tech. Rep. AIC FRANCE A 03/16, 6 pp.
Doviak, R. J., and D. S. Zrnić, 1984: Doppler Radar and Weather Observations. Academic Press, 458 pp., https://doi.org/10.1016/B978-0-12-221420-2.50005-3.
Frehlich, R., 1992: Laser scintillation measurements of the temperature spectrum in the atmospheric surface layer. J. Atmos. Sci., 49, 1494–1509, https://doi.org/10.1175/1520-0469(1992)049<1494:LSMOTT>2.0.CO;2.
Gilson, W., 1992: Radar measurements of aircraft wakes. Massachusetts Institute of Technology Lincoln Laboratory Tech. Rep. AAW-11, 21 pp.
Hill, R. J., 1978a: Models of the scalar spectrum for turbulent advection. J. Fluid Mech., 88, 541–562, https://doi.org/10.1017/S002211207800227X.
Hill, R. J., 1978b: Spectra of fluctuations in refractivity, temperature, humidity, and the temperature-humidity cospectrum in the inertial and dissipation ranges. Radio Sci., 13, 953–961, https://doi.org/10.1029/RS013i006p00953.
Hill, R. J., and S. F. Clifford, 1978: Modified spectrum of atmospheric temperature fluctuations and its application to optical propagation. J. Opt. Soc. Amer., 68, 892–899, https://doi.org/10.1364/JOSA.68.000892.
Jeanmart, H., 2002: Investigation of novel approaches and models for large-eddy simulation of turbulent flows. Doctoral thesis, Ecole Polytechnique de Louvain.
Jeanmart, H., and G. Winckelmans, 2007: Investigation of eddy-viscosity models modified using discrete filters: A simplified “regularized variational multiscale model” and an “enhanced field model.” Phys. Fluids, 19, 055110, https://doi.org/10.1063/1.2728935.
Kovalev, D., and D. Vanhoenacker-Janvier, 2017: Validation of electromagnetic wind radar simulator based on LES with scanning X-band radar measurements and meteorological data. 38th Conf. on Radar Meteorology, Chicago, IL, Amer. Meteor. Soc., 45, https://ams.confex.com/ams/38RADAR/meetingapp.cgi/Paper/320370.
Kovalev, D., D. Vanhoenacker-Janvier, R. Wilson, and F. Barbaresco, 2016: Electromagnetic wind radar simulator validation using meteorological data and a zenith X-band radar. European Radar Conf., London, United Kingdom, Institute of Electrical and Electronics Engineers, 121–124.
Krause, S., 2003: Aircraft Safety: Accident Investigations, Analyses, and Applications. 2nd ed. McGraw-Hill Education, 483 pp.
Li, J., X. Wang, T. Wang, and C. Shen, 2009: Delaminating quadrature method for multi-dimensional highly oscillatory integrals. Appl. Math. Comput., 209, 327–338, https://doi.org/10.1016/j.amc.2008.12.061.
Li, J., X. Wang, and T. Wang, 2010: Scattering mechanism of aircraft wake vortices generated in clear air. Radar Conf., Washington, DC, Institute of Electrical and Electronics Engineers, 117–122, https://doi.org/10.1109/RADAR.2010.5494642.
Li, J., X. Wang, T. Wang, S. Xiao, and M. Zhu, 2011: On an improved-Levin oscillatory quadrature method. J. Math. Anal. Appl., 380, 467–474, https://doi.org/10.1016/j.jmaa.2011.03.055.
Marshall, R. E., and T. J. Myers, 1996: Wingtip generated wake vortices as radar target. IEEE Aerosp. Electron. Syst. Mag., 11, 27–30, https://doi.org/10.1109/62.544796.
Marshall, R. E., A. Mudukutore, V. L. H. Wissel, and T. Myers, 1997: Three-centimeter Doppler radar observations of wingtip-generated wake vortices in clear air. NASA Tech. Rep. CR-97-206260, 90 pp.
Muschinski, A., and S. M. de Bruyn Kops, 2015: Investigation of Hill’s optical turbulence model by means of direct numerical simulation. J. Opt. Soc. Amer., 32A, 2423–2430, https://doi.org/10.1364/JOSAA.32.002423.
Muschinski, A., P. P. Sullivan, D. B. Wuertz, R. J. Hill, S. A. Cohn, D. H. Lenschow, and R. J. Doviak, 1999: First synthesis of wind-profiler signals on the basis of large-eddy simulation data. Radio Sci., 34, 1437–1459, https://doi.org/10.1029/1999RS900090.
Myers, T. J., W. A. Scales, and R. E. Marshall, 1999: Determination of aircraft wake vortex radar cross section due to coherent Bragg scatter from mixed atmospheric water vapor. Radio Sci., 34, 103–111, https://doi.org/10.1029/98RS02776.
Nespor, J., B. Hudson, R. Stegall, and J. Freedman, 1992: Doppler radar detection of vortex hazard indicators. Proc. Aircraft Wake Vortices Conf., Washington, DC, Federal Aviation Administration, 37.
Noonkester, V. R., and J. H. Richter, 1980: FM-CW radar sensing of the lower atmosphere. Radio Sci., 15, 337–353, https://doi.org/10.1029/RS015i002p00337.
Oude Nijhuis, A. C. P., and Coauthors, 2018: Wind hazard and turbulence monitoring at airports with lidar, radar, and mode-s downlinks: The UFO project. Bull. Amer. Meteor. Soc., 99, 2275–2293, https://doi.org/10.1175/BAMS-D-15-00295.1.
Rooseleer, F., and V. Treve, 2013: RECAT-EU—European proposal for revised wake turbulence categorisation and separation minima on approach and departure. EUROCONTROL Tech. Rep., 32 pp.
Rubin, W. L., 2000: Radar-acoustic detection of aircraft wake vortices. J. Atmos. Oceanic Technol., 17, 1058–1065, https://doi.org/10.1175/1520-0426(2000)017<1058:RADOAW>2.0.CO;2.
Sagaut, P., 2006: Large Eddy Simulations for Incompressible Flows. Springer-Verlag, 558 pp., https://doi.org/10.1007/b137536.
Shariff, K., and A. Wray, 2002: Analysis of the radar reflectivity of aircraft vortex wakes. J. Fluid Mech., 463, 121–161, https://doi.org/10.1017/S0022112002008674.
Shephard, D. J., A. P. Kyte, and C. A. Segura, 1994: Radar wake vortex measurements at F and I band. Colloquium on Radar and Microwave Imaging, London, United Kingdom, Institution of Engineering and Technology, 7/1–7/5.
Tatarskii, V. I., 1971: The Effects of the Turbulent Atmosphere on Wave Propagation. Keter Press, 488 pp.
Taylor, G. I., 1938: The spectrum of turbulence. Proc. Roy. Soc. London, 164A, 476–490, https://doi.org/10.1098/rspa.1938.0032.
Vanhoenacker-Janvier, D., K. Djafri, and F. Barbaresco, 2012: Model for the calculation of the radar cross section of wake vortices of take-off and landing airplanes. Ninth European Radar Conf., Amsterdam, Netherlands, IEEE, 349–352.
Vanhoenacker-Janvier, D., C. Pereira, M. Duponcheel, and A. Oude Nijhuis, 2014: Radar simulator for wind monitoring, UFO deliverable: D2110-2. Université catholique de Louvain–Delft University of Technology.
Williamson, J. H., 1980: Low-storage Runge–Kutta schemes. J. Comput. Phys., 35, 48–56, https://doi.org/10.1016/0021-9991(80)90033-9.