\affiliation

[1]organization=Pprime Institute CNRS, ENSMA, University of Poitiers, addressline=1 Av. Clément Ader, postcode=86360, city=Chasseneuil-du-Poitou, country=France \affiliation[2]organization=Laboratory LIAS - ENSIP, University of Poitiers, addressline=2 rue Pierre Brousse, postcode=86073, city=Poitiers, country=France \affiliation[3]organization=STELLANTIS, Advanced Innovation, addressline=212 Bd Pelletier, city=Carrières-sous-Poissy, postcode=78955, country=France

Large scale response of a vehicle wake to on-road perturbations

(April 2024)
Abstract

The aim of this research work is to analyse the large scale response of a vehicle wake to on-road perturbations by using an instrumented vehicle and a combination of scale one wind tunnel tests, track trials and on road experiments. More precisely, in all these tests, we focus on the analysis of the asymmetry of the pressure distribution at the base. Proper Orthogonal Decomposition (POD) is used. For all cases considered, POD analysis reveals two dominant modes, respectively associated with vertical and horizontal wake large scale reorganisation. More than 50% of the total energy is carried by these two modes and this value increases significantly for on-road tests. Noteworthy, the low-frequency energy content of the temporal coefficients of these modes is significantly higher on-road. Low frequencies (even very low ones) then play a major role, corresponding to a quasi-static perturbation domain of the velocity seen by the vehicle. We show that a quasi-steady exploration of the on-road yaw angle statistical distribution during a wind tunnel test captures phenomena similar to those observed on the road and is therefore interesting to evaluate on-road aerodynamic performances. This also opens perspectives for developing closed loop control strategies aiming to maintain a prescribed wake balance in order to reduce drag experienced on the road.

keywords:
Automotive aerodynamics, Drag Reduction, Wind Tunnel tests, On-Road tests

1 Introduction

The European Environment Agency has assessed the impact of transports on greenhouse gases and state that ”Transport is responsible for a quarter of the EU’s greenhouse gas emissions, with road transport representing the greatest share (72% in 2019).111https://www.eea.europa.eu//publications/transport-and-environment-report-2021 Moreover, it is estimated that passenger cars are still the dominant transportation mean with a share of approximately 80% 222https://www.eea.europa.eu/publications/transport-and-environment-report-2022/transport-and-environment-report/view. These studies highlight the importance of reducing the greenhouse gases emissions of road transportation and in this regards vehicle aerodynamics plays a crucial role. In fact, at highway speeds, approximately 70% of the energy losses can be attributed to aerodynamic forces (Kadijk and Ligterink [1], Hucho and Sovran [2]). In that respect, vehicle manufacturers put a lot of effort in optimizing vehicle’s aerodynamics both with numerical simulations and wind tunnel testing. However, the vehicle’s surroundings are constantly changing in real-life scenarios and the drag coefficient, usually optimised at zero yaw condition, varies continuously.

As yaw angle is concerned, several studies have assessed an on-road average yaw angle distribution (Carlino et al. [3], Garcia de la Cruz et al. [4], Yamashita et al. [5], Stoll and Wiedemann [6]), showing that, on average, this distribution corresponds to a quasi-normal distribution centered around 0° with a standard deviation varying upon the external perturbations of the velocity seen by the vehicle due not only to wind turbulence but also to varying road surroundings such as guardrails, vehicles, bushes and trees among others. Hence, only considering the zero yaw angle drag coefficient translates into an under-prediction of the actual averaged drag coefficient. Howell et al. [7] have addressed this problem by defining a wind averaged drag coefficient, on a WLTP (Worldwide Harmonised Light vehicles Test Procedure) road cycle, using quasi-steady approaches :

CDWC=0.530CD0+0.345CD5+0.130CD10+0.007CD15,subscript𝐶𝐷𝑊𝐶0.530subscriptsubscript𝐶𝐷00.345subscriptsubscript𝐶𝐷50.130subscriptsubscript𝐶𝐷100.007subscriptsubscript𝐶𝐷15C_{DWC}=0.530{C_{D}}_{0}+0.345{C_{D}}_{5}+0.130{C_{D}}_{10}+0.007{C_{D}}_{15},italic_C start_POSTSUBSCRIPT italic_D italic_W italic_C end_POSTSUBSCRIPT = 0.530 italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + 0.345 italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT + 0.130 italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT + 0.007 italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ,

CDisubscriptsubscript𝐶𝐷𝑖{C_{D}}_{i}italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (i=0,5,10,15𝑖0510superscript15i=0,5,10,15^{\circ}italic_i = 0 , 5 , 10 , 15 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT) being the drag coefficient at a given yaw angle and CDWCsubscript𝐶𝐷𝑊𝐶C_{DWC}italic_C start_POSTSUBSCRIPT italic_D italic_W italic_C end_POSTSUBSCRIPT being the averaged drag coefficient based on a typical wind distribution. The authors of this study also specifically highlight the importance of reducing the sensitivity of the aerodynamic loads to the on-road perturbations.

Cooper and Watkins [8] and Watkins and Cooper [9] successfully measured the unsteady flow characteristics seen by a road vehicle. They showed that the mean turbulence intensity is of the order of 5% with integral length scales of the order of tens of meters. Moreover, turbulence seen by the vehicle has a rich spectra over a wide range of frequencies and it has been found that the wind gusts feed the spectrum in the low frequencies. Similar results have been found by Wordley and Saunders [10], Wordley and Saunders [11] and Schröck et al. [12]. The first two studies focus on the effect of four different terrain variations on the turbulence scales and intensities. It is shown that ”the four different terrain overlap in the region of lower turbulence intensity (<5%absentpercent5<5\%< 5 %) and smaller turbulent length scale (<5mabsent5𝑚<5\;m< 5 italic_m)”. The study of (Schröck et al. [12]) emphasizes that, when the wind intensity grows, the energy content of the turbulence spectrum seen by the vehicle shifts to the low frequency region, resulting in about 75% of the energy content at frequencies below 2Hz2𝐻𝑧2\;Hz2 italic_H italic_z.

All these studies show the importance of on-road turbulence on vehicles aerodynamics. As an important follow-up on all these contributions, our objective in this paper is to provide data and to analyze the spatio-temporal distribution of the static pressure at the base of a vehicle submitted to such on road perturbations. In particular, it is interesting to determine if the external perturbations seen by the vehicle result in vertical or horizontal large scale asymmetries of the wake that we expect to contribute significantly to the mean drag experienced by the vehicle. To achieve this objective, we equipped a vehicle with a Prandtl antenna and a Conrad probe at the front as well as 49 pressure taps at the base. The vehicle was tested in a wind tunnel, on a track and a ”RDE-like” route (RDE: Real Driving Emissions). Through the collected data, we analyzed the pressure asymmetry of the wake and compared results between wind tunnel experiments and real-world road testing.

The paper is organised as follows: Section 2 gives information on the measurement systems installed in the car and Section 3 describes the different experimental setup. In Section 4 we focus on the wind tunnel experiments at zero yaw angle. Section 5 will highlight the major results obtained on-road and some comparison with the wind tunnel results will be made. Section 6 provides a statistical comparison between the on-road tests and a test with a dynamically varying yaw angle in the wind tunnel. Concluding remarks are proposed in Section 7.

2 Measurement systems

A Citroën C4 Cactus, which is shown in Figure 1, has been used with some added equipment. The vehicle has been equipped with a Prandtl antenna, a Conrad probe, a GPS system and 49 pressure taps at the base. All the equipment is shown in Figures 2 and 3.

Refer to caption
Figure 1: Front-side and back view of the Citroën C4 Cactus, respectively on the left and right-hand side.
Refer to caption
Figure 2: Measuring equipment installed on the vehicle. (a) Prandtl antenna schematisation. (b) Conrad probe schematisation. (c) GPS system mounted inside the vehicle. (d) Pressure scanner mounted in the trunk of the car.
Refer to caption
Figure 3: Pressure taps location on the vehicle’s base. (a) Physical location of the pressure taps. (b) Position schematisation of the pressure taps.

Both Prandtl (Figure 2 a) and Conrad (Figure 2 b) probes have been constructed internally in a Stellantis’ workshop, mainly following the instructions detailed in Chue [13]. They have been calibrated and tested by Thomann [14]. The Prandtl antenna grants an angular tolerance of ±10plus-or-minussuperscript10\pm 10^{\circ}± 10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT while the Conrad probe has a maximum sensitivity of 60superscript6060^{\circ}60 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT with a linear trend for 15β15superscript15𝛽superscript15-15^{\circ}\leq\beta\leq 15^{\circ}- 15 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ≤ italic_β ≤ 15 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. The probes are respectively linked with a Texxis DPS 50 - 0.5 and a Texxis DPS 50 - 2.5 differential pressure scanners. In Figure 2 (c) we can see the GPS system mounted inside the car. Highlighted in green the VBOX Racelogic which is the GPS system itself while highlighted in red there is the CAN box which gather the signals from the two probes and send them to the VBOX. In Figure 2 (d), the differential pressure scanner Scanivalve MPS4264, mounted in the trunk of the car, is shown. It counts 64 total pressure entries and works within a range of 8” H2Osubscript𝐻2𝑂H_{2}Oitalic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_O. Since it is a differential pressure scanner, this means that the pressure on each of the 64 taps is given by pi=paiprsubscript𝑝𝑖subscriptsubscript𝑝𝑎𝑖subscript𝑝𝑟p_{i}={p_{a}}_{i}-p_{r}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, where paisubscriptsubscript𝑝𝑎𝑖{p_{a}}_{i}italic_p start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the absolute pressure on the ithsubscript𝑖𝑡i_{th}italic_i start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT tap, prsubscript𝑝𝑟p_{r}italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the reference pressure, and pisubscript𝑝𝑖p_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the measured value on the ithsubscript𝑖𝑡i_{th}italic_i start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT tap. In our specific case, the reference pressure is obtained using a pierced aluminum bottle mounted inside the trunk of the car. This allow to always have the static pressure in the trunk of the car while minimising the pressure fluctuations by acting like a low-pass filter. In Figure 3 (a) we can see the physical location of the pressure taps at the base of the car whereas in Figure 3 (b) a schematisation of the pressure tap position is given. In the latter, we can also visualize the location of the point considered as the origin as well as the width W=1150mm𝑊1150𝑚𝑚W=1150\;mmitalic_W = 1150 italic_m italic_m and height H=900mm𝐻900𝑚𝑚H=900\;mmitalic_H = 900 italic_m italic_m of the base. The length of the vinyl tubes connected to the scanner varies considerably. At the base, this length can range from 0.7 meters to 2.6 meters, depending on the pressure tap considered. For each measurement channel and therefore each tube length, a phase and amplitude correction was carried out by convolution with a specifically determined transfer function (see Gravier [15]). To find out the transfer function, the method proposed by Tijdeman and Bergh [16] was used taking into account the geometric properties of the circuit for each channel. In order to properly validate some sensitive information from the physical model derived by these authors, such as for example the radius of the vinyl tubes, a calibration was carried out on a specific test bench for several lengths. The calibration bench also made it possible to verify that, over the frequency range considered here (f < 20Hzabsent20𝐻𝑧<\ 20Hz< 20 italic_H italic_z) and after correction, the phase and module of the signals are accurately corrected.

The acquired pressure data is then post-processed to calculate the pressure coefficient:

Cpi=piprQ,subscript𝐶subscript𝑝𝑖subscript𝑝𝑖subscript𝑝𝑟𝑄C_{p_{i}}=\frac{p_{i}-p_{r}}{Q},italic_C start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT = divide start_ARG italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG italic_Q end_ARG , (1)

where pisubscript𝑝𝑖p_{i}italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represents the pressure measured at the ithsubscript𝑖𝑡i_{th}italic_i start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT pressure tap, prsubscript𝑝𝑟p_{r}italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the reference pressure in the vehicle trunk, and Q=12ρU2𝑄12subscript𝜌superscriptsubscript𝑈2Q=\frac{1}{2}\rho_{\infty}U_{\infty}^{2}italic_Q = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ρ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the dynamic pressure, with ρsubscript𝜌\rho_{\infty}italic_ρ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT as the air density and Usubscript𝑈U_{\infty}italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT as the free-stream velocity. During on road tests, Usubscript𝑈U_{\infty}italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the velocity deduced by the GPS system at the time of the measurement.

The following notation will be used for spatial (Adelimited-⟨⟩𝐴\langle A\rangle⟨ italic_A ⟩) and temporal (A¯¯𝐴\overline{A}over¯ start_ARG italic_A end_ARG) averaging, respectively:

A(tm)=1Nn=1NA(xn,tm),delimited-⟨⟩𝐴subscript𝑡𝑚1𝑁superscriptsubscript𝑛1𝑁𝐴subscriptx𝑛subscript𝑡𝑚\langle A\rangle(t_{m})=\frac{1}{N}\sum_{n=1}^{N}A(\textbf{x}_{n},\,t_{m}),⟨ italic_A ⟩ ( italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_A ( x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) ,
A¯(xn)=1Mm=1MA(xn,tm),¯𝐴subscript𝑥𝑛1𝑀superscriptsubscript𝑚1𝑀𝐴subscriptx𝑛subscript𝑡𝑚\overline{A}(x_{n})=\frac{1}{M}\sum_{m=1}^{M}A(\textbf{x}_{n},\,t_{m}),over¯ start_ARG italic_A end_ARG ( italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_M end_ARG ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_A ( x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) ,

where N𝑁Nitalic_N and M𝑀Mitalic_M are the number of pressure taps at the base and the number of time steps, respectively.

However, it is necessary to note that measuring the reference pressure on-road remains a very complicated task because the slightest variation in the vehicle results in a variation in the pressure. Comparisons between the speed data obtained using the Prandtl antenna and those obtained using tachymetric measurements have shown that the reference pressure is not constant. The instantaneous difference between the pressure coefficient of each sensor and the spatial average of the pressure coefficient at the base at the same instant is defined as :

Cp(xn,tm)=Cp(xn,tm)Cp(tm).superscriptsubscript𝐶𝑝subscriptx𝑛subscript𝑡𝑚subscript𝐶𝑝subscriptx𝑛subscript𝑡𝑚delimited-⟨⟩subscript𝐶𝑝subscript𝑡𝑚C_{p}^{*}(\textbf{x}_{n},\,t_{m})=C_{p}(\textbf{x}_{n},\,t_{m})-\langle C_{p}(% t_{m})\rangle.italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) = italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) - ⟨ italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) ⟩ . (2)

Hereafter, we will always use Cp(xn,tm)superscriptsubscript𝐶𝑝subscript𝑥𝑛subscript𝑡𝑚C_{p}^{*}(x_{n},\,t_{m})italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) in order to eliminate any temporal variation in the reference pressure during a test. By using Cp(xn,tm)superscriptsubscript𝐶𝑝subscript𝑥𝑛subscript𝑡𝑚C_{p}^{*}(x_{n},\,t_{m})italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ), we therefore focus on the asymmetry of the wake corresponding to the deviation from the spatial average of the instantaneous pressure distribution at the base.

3 Experimental setup

For this study, we were able to make use of the Stellantis wind tunnel located in Orbassano (10043, Italy) (Stellato and Betti [17]), the Stellantis track of La Ferté-Vidame (28340, France) and a RDE-like route (Yvelines, 78000, France), which are shown in the Figures 4 and 5 respectively.

Stellantis Wind Tunnel

Refer to caption
Figure 4: Schematisation of the Stellantis Wind tunnel in Orbassano (10043, Italy).

From the control room (1), operators can activate the wind tunnel. The test section (2) is 3/4 open, with a variable nozzle section ranging from 22m222superscriptm222\,\text{m}^{2}22 m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to 30.5m230.5superscriptm230.5\,\text{m}^{2}30.5 m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in a plenum measuring 10.5 meters long, 12.2 meters wide, and 10.8 meters high. The turbulence rate at the center of the section is less than 0.5%, and the boundary layer, at X=2.5m𝑋2.5mX=2.5\,\text{m}italic_X = 2.5 m from the exit of the convergent, has a thickness of 5mm5mm5\,\text{mm}5 mm with the boundary layer suction function activated. The maximum test speed is U=215km/hsubscript𝑈215km/hU_{\infty}=215\,\text{km/h}italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 215 km/h with a system of 5 moving belts, one under the car and the remaining four under the wheels to allow their rotation. The test speed for our tests is constant at U=110km/hsubscript𝑈110km/hU_{\infty}=110\,\text{km/h}italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 110 km/h. The car is positioned on a turntable (3) measuring 7 meters in diameter. Beneath the turntable, a six-component balance (4) is installed to measure aerodynamic forces. The vehicle is attached to the balance using four masts, with which we can also adjust the vehicle’s ground clearance, set to the same value as measured with two people inside the vehicle for road tests. Through the heat exchanger (5), we control the temperature to remain constant at T=22C𝑇superscript22CT=22^{\circ}\text{C}italic_T = 22 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT C throughout the tests. The wind is set in motion with a motor-driven fan (6).

On-road routes

Refer to caption
Figure 5: On-road routes visualisation. (a) RDE-like route in the Yvelines (78000, France). (b) Stellantis track of La Ferté-Vidame (28340, France).

The RDE-like route (part (a) of the Figure 5) is a route that partially fullfils to the kilometer distribution in accordance with the new regulatory guidelines for the vehicle homologation process (WLTP). It measures approximately 68 kilometers and lasts about 1 hour. The whole route has, most of the time, allowed vehicle speed above 70km/h70𝑘𝑚70\;km/h70 italic_k italic_m / italic_h. This criterion was chosen because an internal study at Stellantis was conducted to determine from which speed aerodynamic efficiency becomes predominant compared to the rolling resistance of the vehicle (Tuan [18]). This results in the following kilometer distribution: 43%similar-toabsentpercent43\sim 43\%∼ 43 % highway (in green), 46%similar-toabsentpercent46\sim 46\%∼ 46 % rural (in red), and 11%similar-toabsentpercent11\sim 11\%∼ 11 % urban (in blue). Atmospheric pressure information as well as wind speed and orientation were obtained from averages of several weather stations in the area.

Simultaneously, tests were conducted at the Stellantis site in La Ferté-Vidame (28340, France) (part (b) of the Figure 5). This site consists of three tracks: the Outer Track (in green), the Bel Air Track (in red), and the Park Track (in blue). The colors correspond to the permitted speeds on the tracks, so the Outer Track corresponds to a highway route, the Bel Air Track corresponds to a rural route, and the Park Track corresponds to an urban route. We mainly conducted tests on the Outer Track and the Bel Air Track. The Outer Track measures approximately 6.5kmsimilar-toabsent6.5𝑘𝑚\sim 6.5\;km∼ 6.5 italic_k italic_m and includes four main turns. This track allows for a speed constantly over 90km/h90𝑘𝑚90\;km/h90 italic_k italic_m / italic_h throughout the test. The Bel Air Track measures approximately 3.3kmsimilar-toabsent3.3𝑘𝑚\sim 3.3\;km∼ 3.3 italic_k italic_m and includes several short-radius turns, allowing for a speed of 70V9070𝑉9070\leq V\leq 9070 ≤ italic_V ≤ 90. These tracks are connected by connecting ramps (sky-blue dashed lines). The La Ferté-Vidame site also has a weather station capable of measuring wind speed and direction, as well as temperature and atmospheric pressure. All tests conducted on the tracks start at point 0 (in sky-blue), which corresponds to the main entrance of the tracks. The direction of travel is indicated by red triangles. However, each day, the direction of travel is reversed to be able to conduct tests in both directions and ensure that the cars and the track degrade uniformly and not biased by the direction of travel. In our tests, the number of people in the vehicle is always two. There is a driver and the person responsible for verifying the correct progress of the acquisition. The passenger sits in the rear seat where a PC mounting system has been installed to ensure safety requirements. A zero calibration is performed before each test to compensate for any pressure signal drift that may be present. This calibration is performed in a closed building to prevent small disturbances from affecting the process. Acquisition is initiated just before starting the test and stopped once the vehicle is stationary again.

4 Wind tunnel testing

The main objective of the wind tunnel sessions is to characterize the wake of the car in terms of global motions and to analyse the frequency of these motions. Subsequently, the same analysis will be carried out on the road to allow for a comparison of the results.

Proper Orthogonal Decomposition (POD) will be used in that respect. This analysis method was first introduced by Lumley [19] with the aim of decomposing the flow’s turbulence into deterministic functions, each of which can capture a portion of the flow’s turbulent kinetic energy. Several works have summarized this method and explained how to use it (Chatterjee [20], Cordier and Bergmann [21], Taira et al. [22], Weiss [23]).

In our specific case, the principle of POD, calculated here by the direct method after substracting the time average of Cp(xi,t)¯¯subscriptsuperscript𝐶𝑝subscriptx𝑖𝑡\overline{C^{*}_{p}(\textbf{x}_{i},t)}over¯ start_ARG italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) end_ARG, consists of decomposing the vector Cp(xi,t)subscriptsuperscript𝐶𝑝subscriptx𝑖𝑡C^{*}_{p}(\textbf{x}_{i},t)italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) into a set of deterministic functions Φ(k)(x)superscriptΦ𝑘x\Phi^{(k)}(\textbf{x})roman_Φ start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( x ) that are modulated by N temporal coefficients a(k)(t)superscript𝑎𝑘𝑡a^{(k)}(t)italic_a start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( italic_t ) as follows:

Cp(xi,t)=Cp(xi,t)¯+k=1Nmodesa(k)(t)Φ(k)(xi).superscriptsubscript𝐶𝑝subscriptx𝑖𝑡¯subscriptsuperscript𝐶𝑝subscriptx𝑖𝑡superscriptsubscript𝑘1subscript𝑁𝑚𝑜𝑑𝑒𝑠superscript𝑎𝑘𝑡superscriptΦ𝑘subscriptx𝑖C_{p}^{*}(\textbf{x}_{i},t)=\overline{C^{*}_{p}(\textbf{x}_{i},t)}+\sum_{k=1}^% {N_{modes}}a^{(k)}(t)\Phi^{(k)}(\textbf{x}_{i}).italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) = over¯ start_ARG italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) end_ARG + ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_m italic_o italic_d italic_e italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( italic_t ) roman_Φ start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) . (3)

It is important to recall that the functions Φ(k)superscriptΦ𝑘\Phi^{(k)}roman_Φ start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT are orthogonal.

The decomposition is performed using singular value decomposition, that provides the functions Φ(k)superscriptΦ𝑘\Phi^{(k)}roman_Φ start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT, which we will call modes hereafter, as well as the eigenvector 𝝀𝝀\bm{\lambda}bold_italic_λ, which contains the contribution of each mode, in decreasing order, to the total variance. In our study, the number of modes corresponds to the number of pressure sensors at the base, which is 49 modes in total. This means that we have a signal decomposition in the form:

Cp(xi,t)=Cp(xi,t)¯+k=149a(k)(t)Φ(k)(xi),superscriptsubscript𝐶𝑝subscriptx𝑖𝑡¯subscriptsuperscript𝐶𝑝subscriptx𝑖𝑡superscriptsubscript𝑘149superscript𝑎𝑘𝑡superscriptΦ𝑘subscriptx𝑖C_{p}^{*}(\textbf{x}_{i},t)=\overline{C^{*}_{p}(\textbf{x}_{i},t)}+\sum_{k=1}^% {49}a^{(k)}(t)\Phi^{(k)}(\textbf{x}_{i}),\\ italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) = over¯ start_ARG italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t ) end_ARG + ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 49 end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( italic_t ) roman_Φ start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ,

with the total energy defined as follows:

ξ=i=1NtapsCpi2¯=k=1Nmodesλk.𝜉superscriptsubscript𝑖1subscript𝑁𝑡𝑎𝑝𝑠¯superscriptsuperscriptsubscript𝐶subscript𝑝𝑖2superscriptsubscript𝑘1subscript𝑁𝑚𝑜𝑑𝑒𝑠subscript𝜆𝑘\xi=\sum_{i=1}^{N_{taps}}\overline{{C_{p_{i}}^{*}}^{2}}=\sum_{k=1}^{N_{modes}}% \lambda_{k}.italic_ξ = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_t italic_a italic_p italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT over¯ start_ARG italic_C start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_m italic_o italic_d italic_e italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT .

Additionally, thanks to the temporal coefficients of the modes (a(k)(t)superscript𝑎𝑘𝑡a^{(k)}(t)italic_a start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT ( italic_t )), it is possible to analyze the frequency characteristics of the projection of the velocity field onto each mode.

In what follows, we will concentrate on the two first POD modes because they contain a significant contribution of the total variance. We remind that, by using Cp(x,t)superscriptsubscript𝐶𝑝x𝑡C_{p}^{*}(\textbf{x},\,t)italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( x , italic_t ), we focus on the asymmetry of the wake.

Static yaw angle analysis

The first analysis concerns a case with zero yaw angle (β=0𝛽superscript0\beta=0^{\circ}italic_β = 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT). The results are presented in Figure 6.

Refer to caption
Figure 6: Results of the wake analysis for a fixed zero yaw angle in the wind tunnel. Presentation of the distribution of β𝛽\betaitalic_β, the first two modes, the percentage of their contribution to the total energy, as well as spectral analysis of the temporal coefficients associated with the two modes.

The distribution of β𝛽\betaitalic_β shown in the Figure (top-left panel) has a mean μβ=0.3subscript𝜇𝛽superscript0.3\mu_{\beta}=-0.3^{\circ}italic_μ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = - 0.3 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and a standard deviation σβ=0.13subscript𝜎𝛽superscript0.13\sigma_{\beta}=0.13^{\circ}italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = 0.13 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. It is not centered at β=0𝛽superscript0\beta=0^{\circ}italic_β = 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT due to some slight human errors, such as alignment to the symmetry plane of the wind tunnel and alignment of the Conrad probe to the vehicle’s symmetry plane. The latter error remains constant throughout our tests. This distribution of β𝛽\betaitalic_β corresponds to typical cases analyzed by automotive manufacturers, i.e., a nearly constant yaw angle throughout the test. Under these conditions, we performed the POD analysis. The energy contribution of the 49 modes is presented in the bottom-left panel of the figure. We can observe that the first two modes contribute significantly to the total wake energy (approximately 45%percent4545\%45 %), so we will analyze these two modes, which are presented in the top-right panel of the figure. It is interesting to note that these modes have well-defined footprints. For the first mode, we can associate this footprint with a vertical tilting of the wake. In contrast, the footprint of the second mode reveals a horizontal tilting. For this configuration, it is relevant to analyze the characteristic frequencies of these two asymmetries. To do this, we calculated the spectra of the two corresponding temporal coefficients with a frequency resolution Δf=5.102HzΔ𝑓superscript5.102𝐻𝑧\Delta f=5.10^{-2}\;Hzroman_Δ italic_f = 5.10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_H italic_z. This frequency resolution corresponds to a division of the signal into 53 blocks, which will be the minimum number of blocks used subsequently. In the bottom-right panel of the figure, a pre-multiplied spectrum is presented showing the spectra of the two temporal coefficients. Regarding the frequencies, we chose to represent dimensionless frequencies to compare the different tests performed. For this purpose, we used the Strouhal number (St=f.H/Vformulae-sequence𝑆𝑡𝑓𝐻𝑉St=f.H/Vitalic_S italic_t = italic_f . italic_H / italic_V), which is the product of the frequency and the advective time scale (H/V𝐻𝑉H/Vitalic_H / italic_V). Note that the frequency range f<20Hz𝑓20𝐻𝑧f<20\;Hzitalic_f < 20 italic_H italic_z discussed in section 2 corresponds here to a Strouhal number of St< 0.6𝑆𝑡0.6St<\ 0.6italic_S italic_t < 0.6, which means that phases and amplitudes are accurately corrected for all taps in the relevant frequency range. Moreover, from measurements performed with reduced scale models in a wind tunnel (Haffner et al. [24]) we do not expect significant higher frequency contributions for large scale events. We observe a large band distribution of energy for both a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT, with a more defined peak at St101𝑆𝑡superscript101St\approx 10^{-1}italic_S italic_t ≈ 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT. As a measure of the important role of low frequencies we can observe that 62% and 63% of the energy of the mode for the first and second coefficients, respectively, are contained for Strouhal values less than 101superscript10110^{-1}10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

5 On-road results

The response of the wake of the vehicle to on-road perturbations of the velocity seen by the vehicle is analysed in this section. We conducted numerous tests and a subset is shown in Table 1. The two on-road tests presented hereafter are the two test cases highlighted in bold in the Table. These two tests provide a comprehensive overview of the wake behavior under different conditions, allowing us to draw meaningful conclusions about the vehicle’s wake behaviour when compared to wind tunnel tests at zero yaw angle. The robustness of these results is discussed in detail in Cembalo [25]. As shown in Table 1, the conclusions discussed below apply to all on-road tests with only minor differences on energy content in the POD decomposition or spectral energy density of the temporal coefficients.

Routes Vwind[km/h]subscript𝑉𝑤𝑖𝑛𝑑delimited-[]𝑘𝑚V_{wind}[km/h]italic_V start_POSTSUBSCRIPT italic_w italic_i italic_n italic_d end_POSTSUBSCRIPT [ italic_k italic_m / italic_h ] Owindsubscript𝑂𝑤𝑖𝑛𝑑O_{wind}italic_O start_POSTSUBSCRIPT italic_w italic_i italic_n italic_d end_POSTSUBSCRIPT Patm[hPa]subscript𝑃𝑎𝑡𝑚delimited-[]𝑃𝑎P_{atm}[hPa]italic_P start_POSTSUBSCRIPT italic_a italic_t italic_m end_POSTSUBSCRIPT [ italic_h italic_P italic_a ] μβ[]\mu_{\beta}\,[^{\circ}]italic_μ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT [ start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ] σβ[]\sigma_{\beta}\,[^{\circ}]italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT [ start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ] % (Φ1+Φ2subscriptΦ1subscriptΦ2\Phi_{1}+\Phi_{2}roman_Φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) % a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT/a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (St101𝑆𝑡superscript101St\leq 10^{-1}italic_S italic_t ≤ 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) ξ/ξ0𝜉subscript𝜉0\xi/\xi_{0}italic_ξ / italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
Track 22 North-West 975 -0.45 1.74 62 79 and 79 3.01
Track 25 South-West 975 -0.19 1.68 62 79 and 79 5.52
Track 13 South 974 -0.44 1.52 63 78 and 79 3.57
Outer-Track 14 South 979 0.02 1.37 56 76 and 81 2.71
Track 16 South-West 978 -0.08 1.54 63 77 and 79 2.79
Track 8 South 987 -0.31 1.04 69 74 and 69 3.01
Outer-Track 3 South-Est 988 -0.34 1.24 56 75 and 78 2.3
Outer-Track 6 Est 987 0.01 1.15 58 77 and 77 2.43
Outer-Track 6 North-Est 987 0 1.10 55 75 and 77 2.24
RDE-like 8 South-West 1021 -0.02 1.23 55 79 and 70 3.26
RDE-like 9 South-Est 1021 -0.02 1.07 63 81 and 70 3.66
RDE-like 12 North-Est 1015 -0.54 1.56 57 76 and 67 2.84
Highway 12 North-Est 1015 -0.18 1.63 53 72 and 72 1.98
RDE-like 21 North 1009 -0.18 1.65 55 76 and 72 3.66
RDE-like 25 North-Est 1009 0.1 1.4 55 76 and 70 3.03
RDE-like 3 South-West 1022 0.02 1.41 63 72 and 65 6.22
RDE-like 6 West 1021 0.1 1.54 60 69 and 69 2.85
Table 1: Some statistics to show the robustness of the on-road tests. Hereafter, the two tests in bold will be analysed in detail.

In Figure 7 (a), the results are from a test on the La Ferté-Vidame track, while in part (b), it corresponds to a test on the RDE-like route. Atmospheric pressure information as well as wind speed and orientation for the two tests are detailed in Table 1.

Refer to caption
Figure 7: Results of wake analysis on track and on the RDE-like route. Presentation of the distribution of β𝛽\betaitalic_β, the first two modes, and the percentage contribution to the total energy as well as spectral analysis of the temporal coefficients associated with the two modes. (a) Results for test 08 on August 6, 2021, at the La Ferté-Vidame track. (b) Results for test 04 on July 22, 2021, on the RDE-like route.

Regarding the test conducted at La Ferté-Vidame, the direction of travel was counterclockwise, as indicated by the red triangles in Figure 5. The test route includes approximately 4 laps on the Outer track and 2 laps on the Bel Air track. There is very little traffic on this road closed to public. When compared to RDE test conditions, the surroundings of the road on the Stellantis track is repeated at each lap while it displays more diversity on public roads.

It is interesting to note that the distributions of β𝛽\betaitalic_β are very different to that observed in Figure 6. On the track (part (a) of the Figure 7), the distribution has a mean μβ=0.45subscript𝜇𝛽superscript0.45\mu_{\beta}=-0.45^{\circ}italic_μ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = - 0.45 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and a standard deviation σβ=1.74subscript𝜎𝛽superscript1.74\sigma_{\beta}=1.74^{\circ}italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = 1.74 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, while on the RDE-like route (part (b) of the Figure 7), the distribution has a mean μβ=0.02subscript𝜇𝛽superscript0.02\mu_{\beta}=-0.02^{\circ}italic_μ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = - 0.02 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and a standard deviation σβ=1.07subscript𝜎𝛽superscript1.07\sigma_{\beta}=1.07^{\circ}italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = 1.07 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. On the track, the distribution is not centered at β=0𝛽superscript0\beta=0^{\circ}italic_β = 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, which is consistent with the counterclockwise direction of the laps. Additionally, the wind speed was higher during the track test, which may explain the difference observed in the standard deviation.

POD modes 1 and 2 have a very clear large scale signature corresponding to respectively vertical and horizontal balance of the wake. They are very similar for both on-road tests. They are also qualitatively similar to the main POD modes obtained in the wind tunnel at zero yaw angle. However, some details differ, for example the signature of mode 2 at the bottom part of the base. The contribution of these two modes to the total energy is now 63%. This is a very significant increase when compared to the wind tunnel tests (45% contribution for these two modes).

On the RDE-like route, a slight increase in the energy contained in mode 1 (about 44%) and a slight decrease in mode 2 (about 19%) is observed. However, the sum of the energy contained in the two modes remains almost unchanged. Looking now at the premultiplied spectra of the temporal coefficients associated to modes 1 and 2, an increase of the energy in the low frequency domain is observed for the first mode. For all these tests, 70% to 80% of the energy content is measured for Strouhal numbers lower than 0.1. Note that, for on-road situations, the vehicle velocity V𝑉Vitalic_V used in the computation of the Strouhal number in figure 7 is the average vehicle velocity during the test.

A higher frequency resolution (Δf=5.103HzΔ𝑓superscript5.103𝐻𝑧\Delta f=5.10^{-3}\;Hzroman_Δ italic_f = 5.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_H italic_z) was obtained by using longer blocks and therefore a lower number of blocks for averaging the spectra (Nblocks13subscript𝑁𝑏𝑙𝑜𝑐𝑘𝑠13N_{blocks}\leq 13italic_N start_POSTSUBSCRIPT italic_b italic_l italic_o italic_c italic_k italic_s end_POSTSUBSCRIPT ≤ 13). The results are presented in Figure 8. The presence of a large band contribution for Strouhal numbers lower than 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT is very clear, especially for the first mode corresponding to the vertical wake balance (Figure 7). We observe that very low frequencies (down to St103𝑆𝑡superscript103St\approx 10^{-3}italic_S italic_t ≈ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) contribute to the power spectrum. At the average speed maintained during the track test, the distance traveled by the car during a period T=1/f𝑇1𝑓T=1/fitalic_T = 1 / italic_f can be estimated as dV.T=H/Stformulae-sequence𝑑𝑉𝑇𝐻𝑆𝑡d\approx V.T=H/Stitalic_d ≈ italic_V . italic_T = italic_H / italic_S italic_t which is very large for Strouhal numbers lower than 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. We stress that a quasi steady response of the wake is expected for these low Strouhal numbers. This is an important aspect because these low frequencies do not only come from wind turbulence. They probably also originate from the variation in the flow seen by the car due to a variable environment having a wide range of longitudinal scales along the road. Indeed, driving along road portions having similar characteristics (freely exposed to side winds, protected in a forest, exposed to urban area features, …) will induce low frequency components corresponding to the time taken to travel along these portions. Of course, sharp changes in wind directions occur while entering to - or emerging from - a sheltered portion. However, careful experimental study by Haffner et al. [24] using a model geometry at reduced scale shows that natural or provoked wake switching between right/left asymmetry is a slow process taking place in approximately 30 convective time scales. This corresponds to St1/30𝑆𝑡130St\approx 1/30italic_S italic_t ≈ 1 / 30. The slow reorganisation of the large scale wake therefore acts as a low-pass filter that cuts high frequency disturbances.

Refer to caption
Figure 8: Results of the spectral analysis of the temporal coefficients with a frequency resolution of Δf=5.103[Hz]Δ𝑓superscript5.103delimited-[]𝐻𝑧\Delta f=5.10^{-3}[Hz]roman_Δ italic_f = 5.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT [ italic_H italic_z ] for the track and RDE-like route tests. (a) Coefficients a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT on the track. (b) Coefficients a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT on the RDE-like route.

6 Statistical analysis at varying yaw angle in the wind tunnel

Going back to the wind tunnel, we can try to reproduce the yaw angle distribution observed on the road. A subset of yaw angles of the vehicle is first chosen. One single continuous wind tunnel test of approximately 20 minutes - the total number of sample is 120687 at a sampling rate of 100 Hz - is used for the test shown in Figure 9. During the test, our choice is to impose a variable measurement time for each vehicle yaw angle of the subset in order to have a data set representative of the statistical distribution of yaw angles observed on the road.

Refer to caption
Figure 9: Results of the wake analysis for a varying yaw angle in the wind tunnel. Presentation of the distribution of β𝛽\betaitalic_β, the first two modes and the percentage of their contribution to the total energy. The time averaged value Cp(x,t)¯¯subscriptsuperscript𝐶𝑝x𝑡\overline{C^{*}_{p}(\textbf{x},t)}over¯ start_ARG italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x , italic_t ) end_ARG at the base for the full test is shown on bottom right plot

Modes 1 and 2 are very similar to on road modes. We observe that nearly 63% of the turbulent kinetic energy is contained in the first two modes. This is also very similar to the energy distribution for on-road tests. Moreover, the energy contained in the first mode has almost doubled compared to the static yaw angle test, which seems to be due to the variation of the yaw angle along the tests, similarly to what we could observe on the road. This may be a good indication of the interest for making these quasi steady tests during vehicle development to test the robustness of a vehicle design to yaw angle perturbations.

Refer to caption
Figure 10: Comparison of total variance ξ𝜉\xiitalic_ξ for the different analyzed tests. In sky blue, the wind tunnel case with fixed β𝛽\betaitalic_β. In dark blue, the wind tunnel case with varying β𝛽\betaitalic_β. In yellow and red, the cases on track and on the RDE-like route, respectively on the left and on the right.

The total energy ξ𝜉\xiitalic_ξ measured in the different tests is compared in Figure 10 to that measured in the wind tunnel for a fixed yaw angle. This ratio increases significantly for the test with dynamic variation of β𝛽\betaitalic_β in the wind tunnel, reaching more than double the reference value (ξ/ξ02.5𝜉subscript𝜉02.5\xi/\xi_{0}\approx 2.5italic_ξ / italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 2.5). The total energy further increases on the road. On the track, the difference is similar to the cases with dynamically varying β𝛽\betaitalic_β in the wind tunnel (ξ/ξ03.01𝜉subscript𝜉03.01\xi/\xi_{0}\approx 3.01italic_ξ / italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 3.01), while on the RDE-like route, the total variance increases even more (ξ/ξ03.66𝜉subscript𝜉03.66\xi/\xi_{0}\approx 3.66italic_ξ / italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 3.66) compared to the value observed in the wind tunnel for fixed yaw, which is a very significant difference.

The imposed quasi static variations of the yaw angle are interesting in order to widen the statistical distribution of the projection of instantaneous pressure footprints on the main POD modes that represent the wake vertical and horizontal asymmetries. These modes are very similar to on-road main modes and this is an interesting property for wind tunnel tests to be confirmed on other vehicle geometries.

Refer to caption
Figure 11: Distribution of the temporal coefficients a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT as function of the yaw angle β𝛽\betaitalic_β in the varying yaw angle test. The red dot is the conditional average of the temporal coefficient for each orientation of the vehicle.

Figure 11 shows the distribution of the first and second temporal coefficients as function of the yaw angle. Each group corresponds to a prescribed orientation of the vehicle during the dynamic tests. The red dot is the conditional average of the temporal coefficient for each orientation of the vehicle. We see that the POD decomposition is able to describe a large span of wake states at the base. It is well known that drag increases with yaw angle. For the largest positive or negative yaw angles considered here, the change of sign of a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT corresponds to a change in left/right asymmetry. As a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT is concerned, we see that positive values correspond to large yaw angles while small yaw angles - and therefore minimum drag - are associated statistically to negative values of a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT. The time averaged value Cp(x,t)¯¯subscriptsuperscript𝐶𝑝x𝑡\overline{C^{*}_{p}(\textbf{x},t)}over¯ start_ARG italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x , italic_t ) end_ARG at the base for the full test is also shown in Figure 9. This mean distribution displays approximately a left/right symmetry and a positive pressure gradient (pressure increases when moving upward). Using the POD decomposition a low order representation of Cp(x,t)subscriptsuperscript𝐶𝑝x𝑡C^{*}_{p}(\textbf{x},t)italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( x , italic_t ) can be obtained using the first two modes only. From the signature of the mean distribution, from the spatial distribution of the two dominant modes shown in Figure 9 and from the conditional values of the temporal coefficients discussed above, it is easy to figure out that the wake is evolving from a positive vertical asymmetry at about zero yaw to a horizontal asymmetry at larger yaw. Indeed, at larger yaw, a positive value of a(1)superscript𝑎1a^{(1)}italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT reduces the vertical gradient while a non zero a(2)superscript𝑎2a^{(2)}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT corresponds to horizontal pressure gradients. Considering the similarities with on-road results, we therefore conclude that a low frequency exploration of such wake large scale states is performed during on-road travel. Therefore, even a slow closed loop control able to maintain a prescribed wake balance would be interesting for drag reduction. This approach is proposed for a reduced scale model using actuated flaps located at the back of a model vehicle in Cembalo [25] and Cembalo et al. [26].

7 Key findings and Conclusions

The aim of this research work was to analyse the effect of on-road perturbations on the asymmetry of a vehicle wake. We described the vehicle used for testing in a full-scale wind tunnel, on a track and on a RDE-like route. Additionally, we detailed the measurement systems and experimental setup. To analyze the wake characteristics in the wind tunnel and on the road, Proper Orthogonal Decomposition (POD) of the pressure distribution on the base of the vehicle was employed. The main results are summarized below:

  • Wind tunnel tests with fixed yaw substantially deviate from road conditions, showing lower modal energy content and different wake modal footprints, yet still corresponding to similar large scale vertical or horizontal wake asymmetries. The low-frequency energy content of the temporal coefficients of the mode is significantly different from that observed on-road.

  • Road tests exhibit a nearly normal distribution of yaw angle, with values consistently between ±10plus-or-minussuperscript10\pm 10^{\circ}± 10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and 5β5superscript5𝛽superscript5-5^{\circ}\leq\beta\leq 5^{\circ}- 5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT ≤ italic_β ≤ 5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT for over 95% of the time. POD analysis also reveals two primary modes, associated with vertical and horizontal wake asymmetries, respectively. More than 60% of the total energy is carried by these two modes. Low frequencies (even very low ones down to St103𝑆𝑡superscript103St\approx 10^{-3}italic_S italic_t ≈ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) play a major role, corresponding to a quasi-static perturbation domain. Indeed, over 70% of the energy of the two main modes is contained within frequencies St101𝑆𝑡superscript101St\leq 10^{-1}italic_S italic_t ≤ 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT;

  • A wind tunnel run can be designed by selecting a subset of vehicle yaw and a variable measurement time for each yaw. The large data set obtained is then representative of the statistical distribution of yaw angles observed on the road. This single run captures phenomena similar to those observed on the road. The footprint of the two main modes are very similar. The energy content in the first two modes is higher and closer to that measured on road. These tests also show that multiple large scale wake states contribute to the first two modes. Considering the similarities with on-road results, we conclude that a low frequency large span exploration of such wake states is performed during on-road travel.

Further studies should complete this analysis. It would be interesting to analyse the effect of pitch angle variations on these large scale modes. Moreover, the tests have been conducted on a single vehicle. It would be interesting to see the robustness of these results on other types of vehicle. We also stress that a quasi steady response of the wake is expected for the energy containing low Strouhal numbers observed on the road. Therefore, the use of steady CFD computations for a subset of yaw angles is interesting to evaluate a vehicle on-road characteristics. Moreover, even a slow closed loop control able to maintain a prescribed wake balance would be promising for drag reduction strategies. This is an ongoing work on 2/5 reduced scale models using recursive subspace-based predictive control as proposed in Cembalo [25] and Cembalo et al. [26].

Acknowledgements

The authors would like to warmly thank Jean Charles Boueilh for invaluable support during the experiments, as well as Yann Goraguer for moral and technical assistance.

Funding

This research project was funded by STELLANTIS and Ministry for Higher Education and Research. Agostino Cembalo wishes to acknowledge support from ANRT scholarship.

Competing interest

The authors report no conflict of interest.

References

  • Kadijk and Ligterink [2012] Gerrit Kadijk and Norbert Ligterink. Road load determination of passenger cars. TNO Report: TNO, 01 2012.
  • Hucho and Sovran [1993] W. H. Hucho and G. Sovran. Aerodynamics of road vehicles. Ann. Rev. Fluid. Mech., 25:485–537, 1993.
  • Carlino et al. [2007] G. Carlino, D. Cardano, and A. Cogotti. A new technique to measure the aerodynamic response of passenger cars by a continuous flow yawing. SAE Technical papers, 0148-7191:16, 2007. doi: https://doi.org/10.4271/2007-01-0902.
  • Garcia de la Cruz et al. [2017] J. Garcia de la Cruz, Rowan Brackston, and Jonathan Morrison. Adaptive base-flaps under variable cross-wind. SAE Technical Papers, 2017, 08 2017. doi: 10.4271/2017-01-7000.
  • Yamashita et al. [2017] T. Yamashita, T. Makihara, K. Maeda, and K. Tadakuma. Unsteady aerodynamic response of a vehicle by natural wind generator of a full-scale wind tunnel. SAE International Journal of Passenger Cars - Mechanical Systems, pages 358 – 368, 03 2017.
  • Stoll and Wiedemann [2018] D. Stoll and J. Wiedemann. Active crosswind generation and its effect on the unsteady aerodynamic vehicle properties determined in an open jet wind tunnel. SAE Technical Paper, 11(2018-01-0722), 2018. doi: https://doi.org/10.4271/2018-01-0722.
  • Howell et al. [2017] Jeff Howell, David Forbes, and Martin Passmore. A drag coefficient for application to the wltp driving cycle. Proceedings of the Institution of Mechanical Engineers, Part D: Journal of Automobile Engineering, 231(9):1274–1286, 2017. doi: 10.1177/0954407017704784.
  • Cooper and Watkins [2007] K. Cooper and S. Watkins. The unsteady wind environment of road vehicles, part one: A review of the on-road turbulent wind environment. SAE Technical Paper, 2007-01-1236:20, 2007. doi: https://doi.org/10.4271/2007-01-1236.
  • Watkins and Cooper [2007] S. Watkins and K. Cooper. The unsteady wind environment of road vehicles, part two: Effects on vehicle development and simulation of turbulence. SAE Technical Paper, 2007-01-1237:17, 2007. doi: https://doi.org/10.4271/2007-01-1237.
  • Wordley and Saunders [2008] Scott Wordley and Jeff Saunders. On-road turbulence. SAE International Journal of Passenger Cars: Electronic and Electrical Systems, 1(1):341–360, 2008. ISSN 1946-4614.
  • Wordley and Saunders [2009] Scott Wordley and Jeff Saunders. On-road turbulence: Part 2. SAE International Journal of Passenger Cars: Mechanical Systems, 2(1):111–137, 2009. ISSN 1946-3995. doi: 10.4271/2009-01-0002.
  • Schröck et al. [2007] D Schröck, N Widdecke, and J Wiedemann. On-road wind conditions experienced by a moving vehicle. In FKFS-Konferenz—Progress in Vehicle Aerodynamics, Stuttgart, 2007.
  • Chue [1975] S.H. Chue. Pressure probes for fluid measurement. Progress in Aerospace Sciences, 16(2):147–223, 1975. ISSN 0376-0421. doi: https://doi.org/10.1016/0376-0421(75)90014-7. URL https://www.sciencedirect.com/science/article/pii/0376042175900147.
  • Thomann [2018] L. Thomann. Mesures aérodynamiques en environnement réel. Technical report, Polytech Tours, 2018.
  • Gravier [2023] A. Gravier. Correction de l’influence des tubes sur la mesure de pression instationnaire. Technical report, ISAE-SUPAERO, 2023.
  • Tijdeman and Bergh [1965] H. Tijdeman and H. Bergh. Theoretical and experimental results for the dynamic response of pressure measuring systems by h.bergh and h.tijdeman. Technical report, NATIONAL AERO AND ASTRONAUTICAL RESEARCH INSTITUT AMSTERDAM, 01 1965.
  • Stellato and Betti [2018] M. Stellato and L. Betti. Fca full scale wind tunnel: Wltp and coast down test performed with wind tunnel method. SAE Technical Paper, 37(2018-37-0017), 2018. doi: doi:10.4271/2018-37-0017.
  • Tuan [2023] A. F. Da Silva Tuan. Adaptative automotive aerodynamics. Technical report, ISAE-SUPAERO, 2023.
  • Lumley [1967] J. L. Lumley. The structure of inhomogeneous turbulent flows. Atmospheric turbulence and radio wave propagation, pages 166–178, 1967.
  • Chatterjee [2000] Anindya Chatterjee. An introduction to the proper orthogonal decomposition. Current Science, 78(7):808–817, 2000. ISSN 00113891. URL http://www.jstor.org/stable/24103957.
  • Cordier and Bergmann [2008] L. Cordier and Michel Bergmann. Proper Orthogonal Decomposition: an overview. In Lecture series 2002-04, 2003-03 and 2008-01 on post-processing of experimental and numerical data, Von Karman Institute for Fluid Dynamics, 2008., page 46 pages. VKI, 2008. URL https://hal.science/hal-00417819.
  • Taira et al. [2017] K. Taira, S.L. Brunton, S.T.M. Dawson, C. W. Rowley, T. Colonius, B.J. McKeon, O.T. Schmidt, S. Gordeyev, V. Theofilis, and L.S. Ukeiley. Modal analysis of fluid flows: An overview. AIAA Journal, 55(12):4013 – 4041, 2017. doi: DOI:10.2514/1.J056060.
  • Weiss [2019] J. Weiss. A tutorial on the proper orthogonal decomposition. American Institute of Aeronautics and Astonautics, pages 1–21, 2019. doi: https://doi.org/10.2514/6.2019-3333.
  • Haffner et al. [2020] Y. Haffner, J. Borée, A. Spohn, and T. Castelain. Mechanics of bluff body drag reduction during transient near wake reversals. J. Fluid Mech., 894:A14, 2020.
  • Cembalo [2024] A. Cembalo. Amélioration de l’efficacité énergétique des voitures : Contrôle actif de la traînée aérodynamique en conditions amont variables. PhD thesis, École Nationale Supérieure de Mécanique et d’Aérotechnique, 2024.
  • Cembalo et al. [2024] A. Cembalo, Patrick Coirault, Jacques Borée, Clément Dumand, and Guillaume Mercère. Active control of road vehicle’s drag for varying upstream flow conditions using a recursive subspace based predictive control methodology. Submitted to Control Engineering Practice, 2024.