Você está na página 1de 14

Introduction of a New Realistic Generic Car Model for Aerodynamic Investigations

Angelina I. Heft, Thomas Indinger and Nikolaus A. Adams


Technische Universitt Mnchen

2012-01-0168
Published 04/16/2012

Copyright 2012 SAE International doi:10.4271/2012-01-0168

ABSTRACT
State of the art aerodynamic research of vehicles often employs strongly simplified car models, such as the Ahmed and the SAE body, to gain general insights. As these models exhibit a high degree of abstraction, the obtained results can only partly be used for the aerodynamic optimization of production vehicles. Aerodynamic research performed on specific vehicles is on the other hand often limited due to their short life span and restricted access. A new realistic generic car model for aerodynamic research - the DrivAer body - is therefore proposed to close this gap. This paper focuses on the development of the model and the first experimental results, namely force and pressure measurements of the different configurations. The experiments were performed in the recently updated Wind Tunnel A of the Institute of Aerodynamics and Fluid Mechanics at the Technische Universitt Mnchen.

exist (see G.M. Le Good [2]): the use of strongly simplified car models and that of production vehicles. Generic car models, such as the SAE model and the Ahmed body, make it easy to relate the observed phenomena to specific areas and thus help to understand basic flow structures. At the same time, more complex flow phenomena, e.g. at the underbody and the wheels/wheelhouses, cannot be reproduced due to the oversimplification of these geometries. On the other hand, it is usually not feasible to investigate these phenomena on a specific production vehicle, as, due to its short life span and restricted access, typically little validation data is available. Recognizing the need for a model combining the strengths of both approaches, various more or less generic models, such as the VW reference car and the MIRA reference car, have been proposed (cf. G.M. Le Good [2]). However, while these reference cars mark a step in the right direction, these models are still too generic to completely understand the complex phenomena occurring at realistic vehicles. To close this gap, the Institute of Aerodynamics and Fluid Mechanics of the Technische Universitt Mnchen (TUM), in cooperation with two major car companies, the Audi AG and the BMW Group, therefore, proposes a new realistic generic car model. The body is based on two typical medium-class vehicles and includes three interchangeable tops and two different underbody geometries to allow for a high universality. To encourage the use of the DrivAer model in independent research projects, the geometry and a comprehensive database with both numerical and experimental results will be published on the website of the institute1.

INTRODUCTION
Due to growing customer consciousness and various national and international agreements, the reduction of CO2 emissions has become increasingly important. As the car sector is one of the big contributors to the overall CO2 emission, it is necessary to lower the fuel consumption of contemporary cars. Aerodynamic optimization of cars still offers big saving opportunities, especially as its importance increases with the use of recuperation systems (see Hucho [1]). To further optimize the car geometry, it is important to understand the occurring aerodynamic phenomena. So far two basic approaches to investigate these aspects of the flow

1http://www.aer.mw.tum.de/en/research-groups/automotive/drivaer

The aim of this paper is to present this new geometry to a broad audience and to provide first experimental results. In the beginning the necessity of a new realistic generic car geometry will be discussed, followed by a short summary of the development of the DrivAer model. In the next section, the experimental wind tunnel setup will be presented and the general approach and the data processing will be explained. This passage will be followed by the discussion of the experimental results categorized into force and pressure measurements. Concluding the paper, the experimental results will be summarized and a short outlook on future investigations of the DrivAer model will be given.

Figure 1. Dimensions of the SAE model.

DEVELOPMENT OF THE DRIVAER MODEL BACKGROUND


The flow around a bluff body moving in proximity of a static ground is governed by the interaction with the ground and highly turbulent separation and reattachment. Both the experimental reproduction of the relative movement between vehicle and ground and unsteady investigations are highly complex and, therefore, associated with high experimental effort. Yet, to be able to further optimize road vehicles it is necessary to completely understand these phenomena. The improvement of wind tunnel facilities, especially as to the introduction of ground effect simulation through moving belts, allows a more precise experimental investigation of time-accurate flow content (see Janssen [3], Cogotti [4]). Another important advance is the development of smaller and more accurate pressure transducers that can easily be fitted into the surface of wind tunnel models. Categorizing the existing aerodynamic approaches in recent papers on the subject of vehicle aerodynamics, G.M. Le Good [2] identifies two main classes: the investigations performed on strongly simplified generic models and those carried out on real production cars. Commonly, especially time-accurate investigations resort to the use of strongly simplified bodies, such as the SAE body, as presented by Cogotti [5], see Figure 1, and the Ahmed body, as described by Ahmed [6], see Figure 2. These models offer the advantage of a reasonable computational and experimental effort compared to production vehicles and the possibility to examine the flow effects of different parts of the vehicle with limited interference effects. Furthermore, in contrast to specific production vehicles, the simplified models offer a broad spectrum of both numerical and experimental validation data and are accordingly well suited for validation purposes. Figure 2. Dimensions of the Ahmed body. On the other hand, as their shapes are very unlike actual car geometries, these insights cannot be readily applied in the development of production vehicles. Complex areas of the car geometry, such as the A- and C-pillars, the highly curved rear end, and the wheelhouse region, are especially impaired. Therefore, during the actual optimization process, often real production car geometries are employed. As these are usually only accessible to a limited group of people, they are rarely featured in more than one published work and, thus, cannot serve for validation purposes. The gap between these two approaches makes a new realistic generic car model that combines the advantages of both model types desirable. While several attempts to satisfy this need have been made, for example by introducing the MIRA reference car or the VW reference car (see G.M. Le Good [2]), the authors of this paper think that these models are still too generic to allow for detailed investigations of complicated flow phenomena. To this end, the DrivAer model is proposed and will be - along with numerical and experimental results made available to the public. The computer-aided design (CAD) geometry will be published on the homepage of the Institute of Aerodynamics and Fluid Mechanics at the TUM and independent experimental and numerical studies using the geometry are strongly encouraged.

DEVELOPMENT
The geometry of the DrivAer model is based on the geometries of two medium sized cars, the Audi A4 and the BMW 3 Series (see Heft et al. [7]). Audi AG and the BMW Group generously provided the CAD data of the different configurations of the original vehicles. The original CAD surfaces were simplified and approximated by characteristic

curves. In the next step, the curves of both original cars were merged to generate the new CAD geometry. Figure 3 shows a sketch of the fastback configuration of the 1:2.5 DrivAer model measured in the Wind Tunnel A of the Institute of Aerodynamics and Fluid Mechanics.

Figure 4. DrivAer body with different tops.

Figure 3. Main Dimensions of the 1:2.5 DrivAer Model. The flow phenomena at the backlight of contemporary production vehicles can be divided into three main groups with distinct aerodynamic behavior (see Hucho [8]): estate back vehicles, fastback vehicles, and notchback vehicles. While the flow over an estate back vehicle detaches at the end of the roof thus creating a big wake region, the flow over a fastback or a notchback vehicle is strongly influenced by the angle of inclination of the backlight. For fastback vehicles the strong vortices emanating from the C-pillars typically induce a downwash region on the backlight, thus, inhibiting the detachment of the flow at the end of the roof or forcing the flow to reattach on the backlight depending on the angle of inclination. These C-pillar vortices are less pronounced for notchback vehicles. Therefore, the flow detaches at the end of the roof at smaller angles of inclination, the reattachment depends, amongst other things, on the length and height of the trunk. To allow for a thorough investigation of these different behaviors, the DrivAer model was developed as a modular concept that comprises three interchangeable tops, as shown in Figure 4 - just as for example the SAE body and the MIRA reference car (see G.M. Le Good [2]). Furthermore, as shown in Figure 5, two different underbody geometries are provided: a smooth underbody for symmetrical investigations and a detailed underbody based on the simplified underbody geometry of the Audi A4.

(a). Detailed Underbody

(b). Smooth Underbody Figure 5. Different underbody configurations: (a) detailed underbody and (b) smooth underbody. The results presented in this paper were obtained using the mock-up configuration of a scaled 1:2.5 DrivAer model, i.e. without considering a cooling flow. However, future investigations will concentrate on cooling configurations for both conventional and electric cars.

EXPERIMENTAL SETUP
The DrivAer model was measured in the recently updated Wind Tunnel A of the Institute of Aerodynamics and Fluid Mechanics at the Technische Universitt Mnchen, a horizontal Gttingen type wind tunnel. The open test section of the Wind Tunnel A has a length LT=4.8m, the nozzle has a height of HN=1.8m and a width of WN=2.4m. Vortex generators are installed at the nozzle exit to reduce the pressure fluctuations induced by the developing shear layers. To allow for ground simulation (GS) with rotating wheels, the wind tunnel has been equipped with a moving belt system (see Mack et al. [9]). As the moving belt lies 60mm higher than the static ground configuration, the effective nozzle height is reduced to HN,eff=1.74m. Especially for the optimization of underbody geometries and the wheelhouse region, it is essential to simulate the relative motion between the vehicle and the ground [3]. The vehicle body is held from above by a central strut while the wheels are supported separately by four horizontal struts from outside of the test section. In the measured configuration, there is no physical connection between the body and the wheels. The model is placed over a polyester-based belt of 1.39m width and a distance of 4.53m. The basic configuration can be seen in Figure 6.

into account. To obtain physically accurate results, the blockage ratio

(1) should be as small as possible (see Hucho [8]). At the same time, it is desirable to satisfy the Reynolds number similarity to ensure the physical similarity of the flow structures:

(2) If measuring a 1:2.5 model, the free stream wind speed should be 2.5 times higher than for the 1:1 vehicle. In vehicle aerodynamics it is common to perform the measurements at the wind speed u=140km/h which would correspond to a necessary free stream velocity of almost u=100 m/s for a 1:2.5 model. On the other hand, it has been observed that the drag coefficient reaches a relatively stable level for higher Reynolds numbers. At the beginning of the DrivAer measurement cycle, therefore, it has to be verified that the drag coefficient reaches a constant level for the chosen Reynolds number. To that end, the drag coefficient at different Reynolds numbers will be examined. Taking the dimensions of the Wind Tunnel A and its capacity into consideration, the chosen 1:2.5 model with a blockage ratio =8% seems to be an adequate compromise between the two requirements. To facilitate the instrumentation of the wind tunnel model with pressure taps, the top part of the model was laminated, while the underbody geometry was cut out of high-density foam. The force measurements were obtained using a main internal 6-component force balance that was placed between the top strut and the model and four separate 1-component force balances attached to the wheels. The forces of all balances were added up and averaged over three measurement intervals of 10s each.

Figure 6. Picture of the experimental setup in the Wind Tunnel A. The moving belt system can operate up to a velocity of 50 m/s. The boundary layer is reduced using a passive boundary layer scoop. For a more detailed description of the wind tunnel setup, its characteristics, such as the static pressure distribution and the boundary layer profile, be kindly referred to Mack et al. [9]. The decision on the size of the wind tunnel model is based on various factors. On the one hand, it is important to take the effects of wind tunnel blockage on the aerodynamic results

The time-averaged pressure measurements were conducted using a multiport pressure measurement system. The system has a full scale range FS=17kPa and an overall accuracy of 0.15%FS and is connected to the surface pressure taps through flexible tubing (see Vogel [10]). Up to 192 ports can be measured successively in one measurement cycle. For the time-averaged measurements a sampling rate of 20 Hz and an averaging period of 10s were chosen. During one measurement cycle, 188 locations distributed over the surface of the model were measured. In this paper, only some of the most relevant regions will be discussed, namely the symmetry plane that was equipped with 61 probes (in the configuration with detailed underbody), the z=60mm

plane which runs approximately through the stagnation point with 21 measurement locations (see Figure 7), the windshield with 16 pressure taps (see Figure 21), the side window directly behind the A-pillar where 11 probes were located (see Figure 24), and finally the rear windows of the estate back and the notchback configuration with 32 probes each (see Figures 25 and 26).

measurements were not conducted at fixed velocities, but rather at fixed Reynolds numbers. The Reynolds numbers were determined as the equivalent of 10m/s, 20m/s, 30m/s and 40 m/s in an air-conditioned dry air environment at 20C and at sea level (see Table 1). The chosen reference length lref=1.84m corresponds to the length of the scaled DrivAer model. Table 1. Correlation between the Reynolds number and the velocity at standard conditions.

Figure 7. Distribution of pressure taps in the z=60mm plane of the 1:2.5 model. Additionally, time-accurate surface pressures were investigated at the backlight of the fastback model. Areas of periodic detachment and reattachment can be identified with spectral estimates. For this purpose 40 miniature pressure transducers of the type HCL12X5P were placed on the rear slant of the fastback (see Figure 8). These pressure transducers have a pressure range of 12.5mbar and feature a typical error of 0.05%FS and a maximum error of 0.25%FS for combined non-linearity and hysteresis [11]. The instantaneous data was obtained during a sampling interval of 120s at a sampling frequency of 2000Hz. The measurements were filtered with a low pass filter of 1000 Hz.

When working with ground simulation, it is furthermore important to separate the aerodynamic from the frictional forces between the moving belt and the wheels. To isolate the aerodynamic forces, a measurement series with operating moving belt and without wind was conducted and a quadratic equation for the frictional forces was derived by curve fitting (see Figure 9).

Figure 9. Rolling resistance over the moving belt velocity. To obtain the correct aerodynamic forces, the rolling resistance is calculated for the current moving belt velocity and subtracted from the measured forces. Figure 8. Distribution of pressure taps at the backlight of the fastback top. The measured forces are depicted as non-dimensional force coefficients. The drag coefficient can be calculated as:

APPROACH AND DATA PROCESSING


As the Wind Tunnel A of the Institute of Aerodynamics and Fluid Mechanics is not temperature and humidity controlled, the measuring conditions can change quite drastically between summer and winter. To maintain comparability, the (3) The pressure samples presented throughout this paper are also non-dimensioned, i.e. calculated as the ratio of the

differential pressure measured at the surface p and the pressure in the plenum p and the dynamic pressure:

The examination of the 1:2.5 DrivAer model at the reduced Reynolds number of 4.87E+6 consequently seems to be a valid approach.

(4) For more convenience a simple categorizing system will be used in the figures and their captions to describe the model configurations. This system will be quickly introduced here, for a complete scheme please be referred to the definitions section at the end of this paper. The acronym E_S_woM_wW, for example, describes the estate back vehicle with smooth underbody, without mirrors, and with wheels. The first part of the acronym stands for the chosen rear end geometry (E: estate back, F: fastback, and N: notchback), the second represents the underbody geometry (D: detailed and S: smooth), while the third and fourth group refer to the presence of the mirrors (wM: with mirrors, woM: without mirrors) and the wheels (wW: with wheels, woW: without wheels).

Figure 10. Correlation between the total drag coefficient and the Reynolds number for F_D_wM_wW. The drag coefficient without ground simulation is by trend higher than with ground simulation, although the difference lessens with growing Reynolds numbers.

EXPERIMENTAL RESULTS FORCE MEASUREMENTS


Correlation between the drag coefficient and the Reynolds number
At first, the correlation between the force coefficients and the Reynolds number with and without ground simulation will be discussed. As mentioned before, it is not possible to carry out the experiments at the correct Reynolds number in the wind tunnel facilities at the Institute of Aerodynamics and Fluid Mechanics. Therefore, it is important to demonstrate that the effects of Reynolds number changes are insignificant at the operating Reynolds number, before conducting further measurements. To obtain these results, force measurements at four different Reynolds numbers (see Table 1) were conducted for each configuration. As the results are quite similar for all configurations, they will be discussed exemplarily for the fastback configuration (F_D_wM_wW, see Figure 10). The drag coefficient cx converges with growing Reynolds numbers towards the value cx=0.278 (with moving belt) and cx=0.284 (without moving belt). A further increase in Reynolds number is not expected to change the results significantly.

Drag coefficients of the different configurations.


Table 2 shows a comparison of the values of the drag coefficients for the mock-up configuration of the original vehicles, i.e. with closed cooling ducts, and the DrivAer model. The comparative data was provided by Audi AG [13] and the BMW Group [14-15]. Table 2. Comparison of the drag coefficients of the different configurations of the DrivAer model without ground simulation with the drag values of the mock-up Audi A4 and the BMW 3 series geometries.

Similar trends can be observed for the original vehicles and the DrivAer model. The estate back has the highest drag coefficient, while the difference between the fastback and the notchback configuration is very small. The drag values of the DrivAer model are distinctly higher than the drag coefficients of the original vehicles. This is due to the fact, that the averaged geometries of the DrivAer model have not been in any way optimized aerodynamically. However, they lie well

2As the Audi A5 Sportback has a wider wheel track and wheels with a bigger section width than the DrivAer model, the drag coefficient presented here has been corrected to compensate these
differences.

3At the time of the development, of the DrivAer model, the portfolio of the BMW Group did not include a 3 Series Coup. The geometric data used for the derivation of the DrivAer model
was obtained adapting the BMW 5 Series Coup. Therefore, a comparison of the aerodynamic data would not offer further insights.

in the range of other mid-sized cars. With appropriate optimization measurements lower levels of drag are possible. Figures 11,12,13 show the drag coefficients for the different configurations comparing the results obtained with (white) and without (grey) ground simulation. For the fastback configuration (see Figure 11), we can clearly distinguish between the trends of the experiments with wheels and those with closed wheelhouses. Considering the results of the experiments with wheels, the drag coefficient measured without moving belt is higher (8-15 counts) than in the cases with ground simulation. This tendency coincides with the findings of Elofsson and Bannister [16]. The differences observed for the moving belt might be due to the interaction of the C-pillar vortices with vortices emanating from the rotating rear wheels. With closed wheelhouses, we cannot observe this trend; in fact, the difference between ground simulation and no ground simulation seems to be negligible (approximately 1 count). If we compare the configuration with detailed underbody and smooth underbody, we see that the underbody geometry accounts for a significant difference of 32-34 counts (with GS), while the mirrors make a difference for 14-16 counts (with GS). Removing the wheels shows a significant reduction of the total drag coefficient of 102 counts (with GS).

The differences induced by the underbody geometry account for 30-31 counts, the mirrors add up to 14-15 counts and the wheels amount to 96-99 counts.

Figure 12. Drag coefficient for different configurations of notchback top. For the estate back geometry, the differences of the drag coefficient between the experiments with and without ground simulation are a lot smaller than those observed for the fastback and the notchback configuration (see Figure 13). This could indicate that the detachment at the rear of the estate back geometry is determined only by the form of the rear end itself and, thus, not that susceptible to slightly changed flow structures. A different behavior between configurations with and without wheels cannot be seen. The difference in drag coefficient between measurements with and without moving belt lies in a range of 0-4 counts which cannot be seen as aerodynamically significant. The underbody geometry accounts for 27 counts, the mirrors make a difference of 12 counts while the wheels only add up to 83-85 counts.

Figure 11. Drag coefficient for different configurations of fastback top. The notchback configuration exhibits similar behavior to the fastback configuration with a clear distinction between the cases with closed wheelhouses where the drag coefficient is equal or rather slightly higher for the cases with ground simulation, and the cases with wheels where the moving belt lessens the drag coefficient by 9-12 counts (see Figure 12). As the rear end geometry is quite similar to that of the fastback model, similar phenomena should be responsible for the structure of the flow.

Figure 13. Drag coefficient for different configurations of Estate back top.

PRESSURE MEASUREMENTS
Pressure coefficient distribution in selected planes
In the following section, the distribution of the pressure coefficient in different planes will be discussed. If not noted differently, the shown distributions correspond to the measurements of the detailed underbody with mirrors and with wheels. Comparison of the different rear end and underbody configurations Figure 14 shows the distribution of the pressure coefficient in the symmetry plane at the top of the vehicle for the different rear end forms (F_D_wM_wW, E_D_wM_wW, and N_D_wM_wW) as a function of the x coordinate. The measurements were performed using ground simulation. The area formed by the x=0 axis and the cp-curve can be used as a measure for the vertical lift forces acting on the upper and lower surface of the car (see Ahmed et al. [16]). The values of the three configurations coincide well at the front of the vehicles. Starting at the exchangeable top, significant differences occur.

like the fastback geometry up to approximately x=700mm where the flow accelerates again including an additional pressure loss, approximately at the height of the rear wheels, followed by a steep pressure recovery over the backlight and a small drop at the end of the vehicle. The pressure distribution of the fastback configuration lies between that of the estate back and the notchback configuration. Both the fastback and the notchback configuration almost reach ambient pressure behind the vehicle. The pressure distribution in the midplane of the detailed underbody is very similar for all three configurations (see Figure 15). Beginning approximately at the rear wheels (x=1100mm), the estate back configuration recovers less pressure than the other two configurations and, therefore, does not meet the ambient pressure behind the vehicle. Between the fastback and the notchback configuration virtually no differences exist.

Figure 15. Distribution of the pressure coefficient in the y=0mm plane at the bottom of the vehicle for the Estate back (E_D_wM_wW), the Notchback (N_D_wM_wW) and the Fastback (F_D_wM_wW) configuration using ground simulation. Figure 14. Distribution of the pressure coefficient in the y=0mm plane at the top of the vehicle for the Estate back (E_D_wM_wW), the Notchback (N_D_wM_wW) and the Fastback (F_D_wM_wW) configuration using ground simulation. From the stagnation point at the front of the vehicle, the pressure drops swiftly as the flow accelerates over the radiator grill and gains steadily over the bonnet of the car. As the flow is backed up at the junction between bonnet and windshield, the pressure coefficient reaches a local maximum after which the pressure decreases again over the windshield. Depending on the top fixated on the body, a different pressure recovery behavior can be observed. While the estate back recovers the pressure faster, the flow detaches at the edge between the roof and the backlight at approximately x=1300mm and, therefore, does not reach the ambient pressure behind the car. The notchback configuration behaves After a strong acceleration of the flow at the front end, the pressure begins to recover and reaches a local level of high pressure at the height of the mirrors. At the height of the rear wheels, another short pressure drop followed by a steep pressure recovery at the end of the vehicle takes place. Just as the behavior at the bottom of the vehicle, the pressure distribution in the horizontal plane through z=60mm (location marked by the black line) does not show great discrepancies between the different configurations. Again the fastback and the notchback version are almost identical (see Figure 16). After the acceleration at the radiator grill, the pressure rises steadily over the side of the vehicle. The flow detaches both at the front and the rear wheels. As the pressure recovers behind the wheels, it can be assumed that the flow reattaches at both locations.

pressure at the base can be partly responsible for the lower drag coefficient.

Figure 16. Distribution of the pressure coefficient in the z=60mm plane for the Estate back (E_D_wM_wW), the Notchback (N_D_wM_wW) and the Fastback (F_D_wM_wW) configuration using ground simulation. Figure 17 shows the influence of the choice of underbody geometry on the pressure distribution at the top of the fastback configuration. While the overall distribution does not change significantly, the pressure drop at the front is more pronounced for the smooth underbody. The differences at the rear end are most likely due to measurement errors. This will be verified during the next measurement cycle.

Figure 18. Influence of ground simulation on the distribution of the pressure coefficient in the y=0mm plane at the top of the fastback vehicle with a smooth underbody (F_S_wM_wW). The different underbody geometries are only slightly influenced by the moving belt, the flow is accelerated more strongly and the cp- values are therefore slightly lower, whereas the detailed underbody is influenced more strongly than the smooth version (see Figure 19). The smooth underbody only shows small deviations in the wheel area. Over the detailed underbody, the flow accelerates at the front wheels and recovers the pressure more slowly with ground simulation; the deviations at the rear wheels are smaller. For the smooth geometry the pressure coefficient is negative over the whole underbody, the underbody works as a diffusor generating negative lift. The detailed underbody on the other hand shows a short region of positive pressure between the wheels.

Figure 17. Influence of the underbody geometry on the distribution of the pressure coefficient in the y=0mm plane at the top of the fastback vehicle (F_S_wM_wW and F_D_wM_wW) using ground simulation. Influence of ground simulation on the distribution of the pressure coefficient Furthermore, the influence of the moving belt on the development of the pressure coefficient in the selected planes was investigated. As the tendencies for the three configurations are very similar, only the fastback configuration (FS_wM_wW) will be discussed further (Figures 18,19,20). On the upper part of the vehicle, virtually no differences between the two cases exist (see Figure 18). The only deviations appear at the rear end where the pressure is slightly higher for the moving belt case. The slightly higher

Figure 19. Influence of ground simulation on the distribution of the pressure coefficient in the y=0mm plane at the underbody of the Fastback vehicle for different underbody configurations (F_S_wM_wW and F_D_wM_wW).

In the horizontal z=60mm plane, the deviations are mainly located behind the wheels where the flow is accelerated through the rotation of the wheels (see Figure 20). Still, these differences are swiftly compensated, so, the overall flow distribution does not change significantly.

total differences are less than 0.5% of the maximum pressure measured on the windshield, Figure 22 (a)). In Figure 22 (b) the differences between the configuration with and without mirrors are compared. The deltas were calculated by subtracting the measured values with mirrors from the values obtained without mirrors. At the bottom of the windshield, the flow accelerates at the height of the side mirrors, as the flow goes around the mirrors, after slightly slowing down before. The changes are about 10% of the maximum pressures measured on the windshield.

Figure 20. Influence of ground simulation on the distribution of the pressure coefficient in the z=60mm plane of the Fastback vehicle with smooth underbody (F_S_wM_wW) with and without ground simulation.

Pressure coefficient distribution on the surface


In the following, the distribution of the pressure coefficient on different surfaces will be examined and the differences between the individual configurations will be discussed. Windshield In Figure 21 the distribution of the non-dimensional pressure coefficient cp on the wind shield is shown. The pressure was measured at 16 locations and the results were afterwards interpolated by a MATLAB routine. At the base of the wind shield, a stagnation area with moderately high pressure can be identified. The flow accelerates towards the roof and towards the A-pillar.

Figure 22. pressure differences on the windshield for (a) measurements with and without ground simulation (F_D_woM_wW) and (b) measurements with and without side mirrors (F_D_wM_wW and F_D_woM_wW). Side Window In Figure 23 the pressure distribution on the side window without mirrors is presented. The whole area is a low pressure zone. At the A-pillar a region of very low pressure is located which indicates the location of the A-pillar vortex. The A-pillar vortex develops close to the root of the A-pillar and detaches before reaching the roof thus creating the local low pressure zone seen in Figure 23. Figure 24 shows the same surface pressure distribution for the configuration with attached mirrors. While the flow is slowed down in the wake of the mirror the A-pillar vortex grows stronger. The pressure further downstream of the mirror is not affected significantly by the mirrors.

Figure 21. Pressure distribution on the windshield (F_D_woM_wW). Comparisons of the distribution of the pressure coefficients with and without moving belt show almost no difference (the

Figure 23. Distribution of the pressure coefficient at the side window without mirror (F_D_woM_wW).

Figure 24. Distribution of the pressure coefficient at the side window with mirror (F_D_wM_wW). Backlight At the entire rear window of the estate back configuration a zone of moderately low pressure has developed with cpvalues ranging from 0.08 to approximately 0.05 (see Figure 25). This distribution shows that the pressure could not be completely compensated before the flow detaches at the backlight. The deltas between the measurements with and without ground simulation and the configuration with and without mirrors are very small and range between 0.005 and +0.005.

Figure 26. Distribution of the pressure coefficient at the rear window of the notchback configuration (N_D_wM_wW). As described before, the fastback top was equipped with 40 time-accurate pressure transducers (see Figure 8). This allows the evaluation of unsteady flow phenomena at the backlight of the fastback top. As it was found that the ground simulation and the presence of the mirrors do not influence the pressure distribution on the backlight significantly, in the following, only the configuration with detailed underbody, without mirrors and with wheels (F_D_woM_wW) with ground simulation will be discussed. To obtain the averaged pressure distribution at the backlight of the fastback configuration, the instantaneous values were averaged. Figure 27 shows that at the C-pillar a large low pressure zone exists which coincides with the C-pillar vortex. The high pressure zone at the lower part of the backlight shows the reattachment of the separation bubble.

Figure 25. Distribution of the pressure coefficient at the rear window of the estate back configuration (E_D_wM_wW). For the notchback configuration, the flow is accelerated over the roof causing a low pressure zone at the beginning of the backlight (see Figure 26). On the backlight itself, the flow begins to slow down which can be seen from the augmenting pressure. The zone with the lowest pressure is located in the top corner close to the C-pillar. This is due to the C-pillar vortex that develops when the flow is accelerated over the Cpillar.

Figure 27. Distribution of the pressure coefficient at the rear window of the fastback configuration (F_D_wM_wW).

The NACA profile of the top strut is believed to influence the flow structures at the backlight significantly, as described by Hetherington and Sims-Williams [18]. To identify interference frequencies and the influence on the drag and lift of the vehicle, further measurements without the moving belt system and the mounting struts will be performed.

interpretation of the instantaneous results will be presented in a supplementary paper. The results show that the flow around the upper part of the model is not significantly altered by the use of ground simulation. On the other hand, the top sting that is used to hold the model distorts the flow at the rear window decisively. To evaluate the magnitude of this error it is planned to conduct further experiments with a fixed ground without the struts. In addition, measurements in different wind tunnels will be performed to estimate the effects of blockage and the Reynolds number dependency.

SUMMARY
In this paper a new realistic generic car geometry - the DrivAer model - has been introduced. The model was tested in the Wind Tunnel A of the Institute of Aerodynamics and Fluid Mechanics at the TUM. It has been demonstrated that the influence of the Reynolds number on the drag coefficient lessens with augmenting Reynolds numbers and that the investigation of the DrivAer body at the reduced Reynolds number of Re=4.87E+6 is therefore feasible. Examining the drag coefficients of the different configurations, two interesting observations can be made: both the fastback and the notchback DrivAer model show a significant impact of the operating moving belt on the results for attached rotating wheels. These differences do not occur when the wheelhouses are closed. Therefore, it is reasonable to deduce that the vortices emanating from the rotating rear wheels interact with the C-pillar vortices at the backlight of the fastback and the notchback geometry. The estate back configuration, on the other hand, is neither for attached wheels nor for closed wheelhouses influenced significantly by ground simulation. Additionally the distributions of the pressure coefficient in different planes were discussed. Again, apart from the backlight itself, the fastback and the notchback geometry show very similar behavior, while the distribution at the estate back geometry shows different results. Ground simulation does not change the pressure distribution at the top of the vehicles. Small deviations exist in the z=60mm plane directly behind the wheels and along the underbody. Furthermore, the authors evaluated the pressure distribution on different parts of the vehicle surface. It has been shown that the moving belt does not influence the distribution of the pressure coefficient at the top of the vehicle significantly. The mirrors have a substantial influence on the distribution on the windshield and the side window as the flow accelerates over the A-pillar and the A-pillar vortex grows stronger.

REFERENCES
1. Hucho, W.-H. (ed.), Reduzierung des Luftwiderstandes voile Wirkung mit regenerativem Bremsen, AutomobilRevue, Nr. 36, 2009. 2. Le Good, G. M., and Garry, K. P., On the Use of Reference Models in Automotive Aerodynamics, SAE Technical Paper 2004-01-1308, 2004, doi: 10.4271/2004-01-1308. 3. Janssen, L.J., Mllenbach, P., and Deutenbach, K.R., Einfluss der Fahrbahnsimulation bei Windkanalversuchen an PKW, J. of Wind Engineering and Industrial Aerodynamics, 22:347-379, 1986, doi: 10.1016/0167-6105(86)90097-8. 4. Cogotti, A., Ground Effect of a Simplified Car Model in Side-Wind and Turbulent Flow, SAE Technical Paper 1999-01-0652, 1999, doi:10.4271/1999-01-0652. 5. Cogotti, A., A Parametric Study of the Ground Effect of a Simplified Car Model, SAE Technical Paper 980031, 1998, doi: 10.4271/980031. 6. Ahmed, S. R., Ramm, G., and Faltin, G., Some Salient Features of the Time-Averaged Ground Vehicle Wake, SAE Technical Paper 840300, 1984, doi: 10.4271/840300. 7. Heft, A.I., Indinger, T., and Adams, N.A., Investigation of Unsteady Flow Structures in the Wake of a Realistic Generic Car Model, AIAA-2011-3669, 2011. 8. Hucho, W.-H. (ed.), Aerodynamik des Automobils, Vieweg+Teubner, Wiesbaden, ISBN 0-13-978-3528039592, 2005. 9. Mack, S., Indinger, T., Adams, N. A., and Unterlechner, P., The Ground-Simulation Upgrade of the TUM Wind Tunnel, SAE Technical Paper 2012-01-0299, 2012, doi: 10.4271/2012-01-0299. 10. Vogel, F., Breitsamter, C., and Adams, N.A., Aerodynamic Investigation on a Helicopter Fuselage, AIAA 2011-3820, 2011.

FUTURE WORK
In a next step the time-accurate measurements will be examined and compared to numerical simulations. The identification of the corresponding vortices and the thorough

11. Sensortechnics GmbH, Puchheim, Germany. HCL Series - Miniature compensated low pressure sensors, 2010. 12. Wickern, G., Drag coefficients of the Audi A4 Saloon, Audi A4 Avant and the Audi A5 Sportback, Audi AG, 2011. 13. EADE European Aerodynamic Data Exchange, BMW AG - 320i: Station Wagon, Chart No. 236, 2008. 14. EADE European Aerodynamic Data Exchange, BMW AG - 320i: Sedan, Chart No. 229, 2008. 15. Elofsson, P. and Bannister, M., Drag Reduction Mechanisms Due to Moving Ground and Wheel Rotation in Passenger Cars, SAE Technical Paper 2002-01-0531, 2002, doi:10.4271/2002-01-0531. 16. Ahmed, S.R., Gawthrope, R.G., and Mackrodt, P.-A., Aerodynamics of Road- and Rail Vehicles, Vehicle System Dynamics, 14(4), 319-392, doi: 10.1080/00423118508968836. 17. Hetherington, B. and Sims-Williams, D. B., Wind Tunnel Support Strut Interference, SAE Technical Paper 2004-01-0806, 2004, doi:10.4271/2004-01-0806.

FS Full Scale Range GS Ground Simulation PSD Power Spectral Density TUM Technische Universitt Mnchen To describe the different configurations of the DrivAer model a simple categorizing system has been derived. In the following short paragraph/passage the system used throughout the paper will be presented. First letter: describes the rear end configuration F Fastback E Estate back N Notchback Second letter: describes the underbody configuration D Detailed Underbody S Smooth Underbody Third block of letters: describes the mirror configuration wM With Mirrors woM Without Mirrors Fourth block of letters: describes the wheel configuration wW

CONTACT INFORMATION
Angelina Heft Institute of Aerodynamics and Fluid Mechanics Technische Universitt Mnchen Boltzmannstr. 15 85748 Garching angelina.heft@aer.mw.tum.de

ACKNOWLEDGMENTS
The authors would like to thank the technical staff at the Institute of Aerodynamics and Fluid Mechanics (TUM), especially W. Ltzenburg and H.-G. Frimberger. Furthermore, we would like to express our thanks to S. Mack for the implementation of the ground simulation in the Wind Tunnel A and the instrumentation of the DrivAer model. We sincerely thank both the Audi AG and the BMW Group for their invaluable support in the development process of the DrivAer body.

DEFINITIONS/ABBREVIATIONS
CAD Computer Aided Design CFD Computational Fluid Dynamics FFT Fast Fourier Transformation

With Wheels

woW Closed Wheel Houses E_S_woM_wW therefore describes the configuration Estate back with smooth underbody, without mirrors and with wheels.

p Pressure p Plenum pressure Re Reynolds number u Wind speed Pressure coefficient Kinematic viscosity Drag coefficient Density Rolling Resistance Wind tunnel blockage ratio

NOMENCLATURE
Aref Reference Area cp

cx

FR

Fx Drag Force HN Height of the Nozzle WN Width of the Nozzle lref Reference Length LT Length of the wind tunnel test section

The Engineering Meetings Board has approved this paper for publication. It has successfully completed SAE's peer review process under the supervision of the session organizer. This process requires a minimum of three (3) reviews by industry experts. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of SAE. ISSN 0148-7191

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely responsible for the content of the paper. SAE Customer Service: Tel: 877-606-7323 (inside USA and Canada) Tel: 724-776-4970 (outside USA) Fax: 724-776-0790 Email: CustomerService@sae.org SAE Web Address: http://www.sae.org Printed in USA

Você também pode gostar