+ All Categories
Home > Documents > The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip...

The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip...

Date post: 15-Sep-2018
Category:
Upload: doannga
View: 214 times
Download: 0 times
Share this document with a friend
18
The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortex Hirofumi Igarashi, Paul A. Durbin and Hui Hu () Department of Aerospace Engineering, Iowa State University, Ames, Iowa, 50011 Email: [email protected] and Scott Waltermire and Joe Wehrmeyer Arnold Engineering Development Center, Arnold AFB, TN Abstract An experimental study was conducted to investigate the behavior of the wing-tip vortex structures generated by a square-tipped, rectangular NACA0012 wing. A Stereoscopic Particle Image Velocimetry (SPIV) system was used to conducted detailed flow field measurements to elucidate the key features of the wing-tip vortex structures in the near field. One of the great advantages of the present SPIV measurements over the classical measurement technique is that the vortex wandering can be removed directly by tracking the center of the wingtip vortex in the instantaneous measurement frames. By tracking the center of the wingtip vortex, the wandering and turbulence in the vortex can be decoupled completely. This method was applied to investigate the effects of the angle of attack of the test wing and wind tunnel wall on the evolution of the wingtip vortex in the near field. In order to decouple the effects of vortex wandering, Devenport et al. (1996) suggested an analytical method to predict the wandering free velocity profile. The velocity profile predicted by the Devenport et al. (1996) method was compared with the SPIV re-centered velocity profile quantitatively. Introduction The behavior of strong, coherent wingtip vortex structures is of great importance to various commercial and military applications. One well-known problem is the hazardous effect of trailing wingtip vortices on flight safety and airport capacity. From a military standpoint, there are many issues associated with the effects of wingtip vortex structures on dynamics of towed vehicles, tail buffeting, and icing arrays. Blade/vortex interaction on helicopter blades can impact performance and cause undesirable noise and vibration. Vortex wandering is the slow side-to-side movement of the wing-tip vortex core. It has been found to be a universal feature of wind-tunnel-generated wing-tip vortex structures. Devenport et al. (1996) developed an analytical technique to correct mean-velocity profiles measured with a fixed hot-wire probe, thereby providing quantitative estimates of wandering amplitude and its contributions to Reynolds stress. This process involves forming a “corrected” mean velocity field from the measured version, using guessed levels of wandering amplitude. The corrected field is then artificially subjected to wandering, described by a bi-variate Gaussian probability density function (PDF) for vortex position, and the resulting Reynolds stresses at the vortex 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition 4 - 7 January 2011, Orlando, Florida AIAA 2011-1165 Copyright © 2011 by the author(s). Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
Transcript
Page 1: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortex

Hirofumi Igarashi, Paul A. Durbin and Hui Hu ( )

Department of Aerospace Engineering, Iowa State University, Ames, Iowa, 50011 Email: [email protected]

and Scott Waltermire and Joe Wehrmeyer

Arnold Engineering Development Center, Arnold AFB, TN

Abstract

An experimental study was conducted to investigate the behavior of the wing-tip vortex structures generated by a square-tipped, rectangular NACA0012 wing. A Stereoscopic Particle Image Velocimetry (SPIV) system was used to conducted detailed flow field measurements to elucidate the key features of the wing-tip vortex structures in the near field. One of the great advantages of the present SPIV measurements over the classical measurement technique is that the vortex wandering can be removed directly by tracking the center of the wingtip vortex in the instantaneous measurement frames. By tracking the center of the wingtip vortex, the wandering and turbulence in the vortex can be decoupled completely. This method was applied to investigate the effects of the angle of attack of the test wing and wind tunnel wall on the evolution of the wingtip vortex in the near field. In order to decouple the effects of vortex wandering, Devenport et al. (1996) suggested an analytical method to predict the wandering free velocity profile. The velocity profile predicted by the Devenport et al. (1996) method was compared with the SPIV re-centered velocity profile quantitatively.

Introduction

The behavior of strong, coherent wingtip vortex structures is of great importance to various commercial and military applications. One well-known problem is the hazardous effect of trailing wingtip vortices on flight safety and airport capacity. From a military standpoint, there are many issues associated with the effects of wingtip vortex structures on dynamics of towed vehicles, tail buffeting, and icing arrays. Blade/vortex interaction on helicopter blades can impact performance and cause undesirable noise and vibration.

Vortex wandering is the slow side-to-side movement of the wing-tip vortex core. It has been found to be a universal feature of wind-tunnel-generated wing-tip vortex structures. Devenport et al. (1996) developed an analytical technique to correct mean-velocity profiles measured with a fixed hot-wire probe, thereby providing quantitative estimates of wandering amplitude and its contributions to Reynolds stress. This process involves forming a “corrected” mean velocity field from the measured version, using guessed levels of wandering amplitude. The corrected field is then artificially subjected to wandering, described by a bi-variate Gaussian probability density function (PDF) for vortex position, and the resulting Reynolds stresses at the vortex

49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition4 - 7 January 2011, Orlando, Florida

AIAA 2011-1165

Copyright © 2011 by the author(s). Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

Page 2: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

center are compared with the actual measured values. The wandering amplitudes are then adjusted iteratively until the calculated and measured values converge. Using the wandering correction theory, they concluded that the vortex core in their experiments was laminar and that velocity fluctuations in the vortex core region were mainly due to wandering. These studies have uncovered useful information, but some of the inconsistencies noted above raise questions. The majority of the previous studies used point-wise flow measurement techniques such as pressure probes, hot-wire anemometry, and laser Doppler anemometer. A common shortcoming of such point-wise measurements is the inability to provide spatial structure of the unsteady vortices. Full field measurements are needed to effectively reveal the transient behavior of the wing-tip vortex structures. Temporally-synchronized and spatially-resolved flow field measurements are highly desirable in order to elucidate underlying physics. Advanced flow diagnostic techniques such as high-resolution stereoscopic Particle Image Velocimetry (SPIV) used in the present study are capable of providing such information.

Based on the flow field measurements with a conventional two-dimensional PIV system, Heyes et al. (2004) evaluated wandering effects by re-centering PIV data. They found that the Devenport et al. (1996) assumption of using a bi-variate normal probability density function could be valid, and their corrections were in good agreement with those predicted by the Devenport et al. (1996) method. They found a 12.5% over-prediction of the core radius and a 6% under-prediction of the peak tangential velocity. The errors were larger for lower angles of attack. It should be noted the PIV measurement results reported in Heyes et al. (2004) was obtained by using a conventional 2-DPIV system. It is well known that a conventional 2-D PIV system is only capable of recording the projection of velocity into the plane of the laser sheet. That means the out-of-plane velocity component is lost while the in-plane components may be affected by an unrecoverable error due to the perspective transformation (Prasad & Adrian, 1993). For the highly three-dimensional flows like wingtip vortex flows, the two-dimensional measurement results may not be able to reveal their three-dimensional features successfully. In the present study, a high-resolution Stereoscopic Particle Image Velocimetry (SPIV) system, which is capable of achieving quantitative measurements of all three components of instantaneous flow velocity vectors, is used to conduct detailed flow field measurements to quantify the transient behavior of the wing-tip vortex generated from a rectangular NACA0012 airfoil.

Experimental Setup and Stereoscopic PIV System The experimental study was conducted in a closed-circuit low-speed wind tunnel located in

the Aerospace Engineering Department of Iowa State University. The tunnel has a test section with a 12.0 × 12.0 in (30 × 30 cm) cross section and the walls of the test section are optically transparent. The tunnel has a contraction section upstream of the test section with honeycombs, screen structures and a cooling system installed ahead of the contraction section to provide uniform low turbulent incoming flow into the test section. The standard deviation of velocity fluctuations at the test section entrance was found to be about 0.8% of the free-stream velocity, measured by using a hotwire anemometer.

Figure 1 shows the schematic of the experimental setup used in the present study. The flow was seeded with 1~5 μm oil droplets. Illumination was provided by a double-pulsed Nd:YAG laser (NewWave Gemini 200) adjusted on the second harmonic and emitting two pulses of 200 mJ at the wavelength of 532 nm with a repetition rate of 5 Hz. The thickness of the laser sheet in

Page 3: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

the measurement region is about 2.0 mm. Two high-resolution 12-bit (1600 x 1200 pixel) CCD camera (PCO1600, CookeCorp) were used to perform stereoscopic PIV image recording. The two CCD cameras were arranged in an angular displacement configuration to get a large overlapped view. In the present study, the angle the angle between the view axial of the two cameras is set to about 60 degrees. For such an arrangement, the size of the stereoscopic PIV measurement window is about 80mm by 60mm. The CCD cameras and double-pulsed Nd:YAG lasers were connected to a workstation (host computer) via a Digital Delay Generator (Berkeley Nucleonics, Model 565), which controlled the timing of the laser illumination and image acquisition.

A general in-situ calibration procedure was conducted in the present study to obtain the mapping functions between the image planes and object planes (Soloff et al. 1997). The mapping function used in the present study was taken to be a multi-dimensional polynomial, which is fourth order for the directions (X and Y directions) parallel to the laser sheet plane and second order for the direction (Z direction) normal to the laser sheet plane. The coefficients of the multi-dimensional polynomial were determined from the calibration images by using a “least square” method. The two-dimensional particle image displacements in each image plane were calculated separately by using a frame to frame cross-correlation technique involving successive frames of patterns of particle images in an interrogation window 32×32 pixels. An effective overlap of 50% of the interrogation windows was employed in PIV image processing. By using the mapping functions obtained by the in-situ calibration and the two-dimensional displacements in the two image planes, all three components of the velocity vectors in the illuminating laser sheet plane were reconstructed (Hu et al. 2002).

Fig. 1: Schematic of the experimental setup

The test model used in the present study was a zero-sweep, untwisted half wing, with a NACA0012 symmetric profile. The wing had a rectangular planform with a half-span b/2=0.1524m and a chord c=0.1016m and a plane tip with sharp edges. Its root was mounted on

Page 4: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

one side of the wind tunnel wall. The angle of attack of the tested wing could be changed using a rotary table mounted outside the wind tunnel with the axis of rotation passing through the airfoil’s aerodynamic centre (quarter-chord point). The wing was mounted through the wind tunnel wall such that its quarter-chord axis was along the centerline of the wind tunnel. Following the work of Bailey & Tavoularis (2008), a boundary layer trip wire was attached on the suction surface of the airfoil, 0.10c away from the leading edge, to induce transition, thus reducing the possibility of separation and sensitivity to free-stream conditions. For the present study, the angle of attack of the tested wing was changed from 0 to 20 degrees (i.e.,

.deg20~0=α ) and the results from 2 to 12 degrees are shown in the present study. During the experiments, the wind-tunnel speed was adjusted from 7.9 m/s ~ 17.3 m/s, which corresponds to a chord Reynolds number .1015.1~1052.0Re 55 ××=C

(a). Without additional walls (b). With additional wall (a=2.0, 4.0 and 5.0in)

Fig. 2: Cross section view of the test section

In order to investigate the effects of the wind tunnel wall on the evolution of the wing tip vortex, additional U-shaped hollow walls are inserted in the test section to reduce the distance between the wing tip and the tip-side wall. The width of the wall, as shown in figure 2, is varied to 2.0, 4.0 and 5.0 inch so that the distance between the wing tip and wall is set to 4.0, 2.0 and 1.0 in respectively. These additional walls are made of optically transparent acrylic plates with the thickness of 1/8 in. This wall is thick enough to stand rigid in the test section; yet, thin enough to cause no noticeable tunnel blockage issues. Since the index of refraction may vary due to the additional walls, in-situ calibration procedures for stereoscopic PIV measurements were performed each time to compensate the optical distortion.

Page 5: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

Experimental Results and Discussions

A. SPIV Method In this section, the method used to extract the wandering free vortex velocity field is

described; hereafter, referred to as the ”SPIV method” to distinguish from the Devenport et al. (1996) correction method. SPIV data presented in this study is the result of 500 instantaneous SPIV frames, taken at the rate of 4 frames per second. Since the SPIV data acquisition rate was far below the frequency of the vortex motion, the velocity fields shown in this study is the time average of 500 instantaneous frames, or about 125 seconds. To extract the wandering free vortex velocity field, it is necessary to keep track of the vortex center at every instantaneous SPIV frame. First, the vortex center is estimated by finding the maximum negative vorticiy around the middle of the measured plane. Then, the first 7.0mm × 7.0mm grid with 8 × 8 nodes are formed with maximum vorticity in the center and the cross stream velocities are interpolated to the minimum values. After determining the minimum cross stream velocity with the 7.0mm × 7.0mm grid, the new grid is formed twice to a 5.0mm × 5.0mm grid with 51 × 51 nodes and further down to a 1.0mm × 1.0mm grid with 101 × 101 nodes. As the program generates new grids, the grid size gets smaller and the grid gets finer. The velocities on the new grid nodes are linearly interpolated from nearby node values in the old grid since the velocity profile is close to linear in the vortex core. Then a 50 to 40mm square velocity field, depending on the case, is extracted with the vortex center at (0, 0) as shown in Figure 3(a). By repeating this method for all the frames, these square velocity fields can be considered as the instantaneous “wandering corrected” velocity fields. Then, the core radius of the wingtip vortex was determined from the measured tangential velocity profiles of the vortex, as shown in Fig. 3(b).

(a).Velocity field with a vortex at center (b). core radius and maximum tangential velocity

Fig. 3: Extracted velocity field and a typical tangential velocity profile.

By tracking the vortex center for all 500 frames, the vortex wandering behavior is quantitatively described with time averaged mean vortex center at (0,0). Figure 4(a) shows the projection of the vortex centers in y-z plane for o0.2=α case and Figures 4(b) and 4(c) show the vortex trajectory in y and z directions, respectively. From the statistical analysis, the amplitude

Page 6: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

of the vortex wandering is described by the standard deviation of the instantaneous vortex center location. The wandering amplitude of the vortex in y-direction is denoted yσ and z-direction is zσ . They are normalized by the core radius, cr , as in Figure 3(b). The correlation of two is shown as ρ to indicate the orientation of the vortex wandering. Figures 5(a) and 5(b) show the histogram of the vortex location which validates one of the assumptions made by Devenport et al. (1996) that the vortex wanders in a Gaussian distribution manner.

-1.0

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1.0

-1.0 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1.0

Y/rc

Z/r c

(a) Vortex center projection in y-z plane

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

0 20 40 60 80 100 120

Time (sec)

Y/r c

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

0 20 40 60 80 100 120

Time (sec)

Z/r c

(b) Vortex center trajectory in y-direction (c) Vortex center trajectory in z-direction

Fig. 4: Vortex center projection and trajectory for o0.2=α ( cx 0.4= , .1015.1Re 5×=C )

Page 7: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(a) y-direction (b) z-direction

Fig. 5: Histogram of vortex center at o0.2=α ( 0.4/ =cx , .1015.1Re 5×=C )

B. The effects of the angle of attack of the test wing Figure 6a shows the angle of attack effect on the wandering amplitude. The circle symbol

indicates the wandering amplitude in y-direction, yσ , and the square symbol indicates the amplitude in z-direction, zσ . As shown in the Figure, the both amplitudes, yσ and zσ , were found to decrease with the angle of attack increases until α ≈ 8O. After α > 8O, the wandering amplitudes was found almost stay constant. This result was found to closely agree with the findings of Devenport et al. (1996). They reported the wandering amplitudes (σ / c) of 0.009, 0.006, 0.006, and 0.005 at α = 2.5, 3.9, 5.0 and 7.5⁰ respectively, where the present study shows that the wandering amplitude of 0.009, 0.005, 0.004, and 0.003 at α = 2.0, 4.0, 6.0 and 8.0 O , respectively. Also, as shown by Devenport et al. (1996), the vortex wanders 0.002c or 0.001c more in y-direction (span-wise direction). This trend can be seen in the present study as the vortex wanders 0.002c or 0.004c more in the span-wise direction. As shown in the Table 1, the wandering is found to have the largest amplitude when the vortex had the lowest peak tangential velocity, Vθmax. When the maximum tangential velocity Vθmax ≈ 2.0m/s at α = 2.0 O, the vortex wanders about 0.9mm in both y and z directions. However, when the angle of attack is increased to 12.0 O, the tangential velocity is increased to Vθmax ≈ 8.60m/s and the vortex wandering was found to be reduced by 70%. For the same flow condition, even thought the wandering is constant at the higher angle of attack, the increase in the tangential velocity will have some effect on decreasing the vortex wandering which agree with the finding by Bailey and Tavoularis (2008).

Figure 6b shows the angle of attack effect on the vortex core radius. Initially the axial-averaged vortex core radius is 5.37mm at o0.2=α . When the angle of attack is increased to o0.4=α , the vortex core radius was found to become smaller to 4.85mm. As the angle of attack gets larger toward o0.12=α , the core radius became bigger and bigger. At o0.6=α , the vortex core radius is 5.0mm and for the larger angle of attack of o0.12=α , the vortex radius was found to increase to 6.19mm. This trend, a small decrease in the vortex core before the increasing core radius at high angle of attack, also can be seen in the study of Gerontakos and

Page 8: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

Lee (2006). By using a sweptback and tapered NACA0015, Gerontakos and Lee (2006) showed the sharp decrease in the core radius up to o0.6=α and the gradual increase as the angle approaches to o0.9=α .

0

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.16

0.18

0.20

0 2 4 6 8 10 12 14

σZ

σY

α (deg)

σ/r c

4.0

4.5

5.0

5.5

6.0

6.5

7.0

0 2 4 6 8 10 12 14

α (deg) r c (m

m)

(a). Wandering amplitude (b). Axial averaged vortex core radius

Fig 6: Angle of attack effect ( 0.4/ =cx , .1015.1Re 5×=C )

Table1: The measured wandering amplitude vs. the estimations by using Devenport’s method.

Figures 7 show the normalized turbulence kinetic energy (T.K.E.) distributions of the wingtip vortex with and without the wandering correction. The left column of the figures show the distributions without the wandering correction and the right column show the distributions with wandering correction. At o0.2=α , the highest normalized T.K.E. value of about 31060.1 −× is observed at the center of the vortex (Figure 7a). After tracking the vortex center in every frame to eliminate the vortex wandering effect, as shown in Figure 7b, the normalized T.K.E. at the vortex center was found to be reduced to less than 4100.1 −× . By comparing the two plots, it can be seen that the higher T.K.E. values in the center of the vortex is mainly due to the vortex wandering rather than turbulence. As the angle of attack increases, the spiral wake becomes larger, while the maximum T.K.E. value in the wake stays about 3100.1 −× . As the angle of attack increasing from o0.8 to o0.12 , the T.K.E. level increases significantly. At o0.8=α , the maximum T.K.E. is about 31060.1 −× at the center but for o0.12=α , the T.K.E. goes above 2100.2 −× and fills most of the vortex core. Even after removing the wandering, the T.K.E. in the vortex core increases from 3100.3 −× to 3100.6 −× (Figure 7e and 7f). This suggests that the wingtip vortex is much more turbulent at higher angles of attack.

Page 9: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(a).

o0.2=α , before wandering correction (b). o0.2=α , after wandering correction

(c). o0.6=α , before wandering correction (d).

o0.6=α , after wandering correction

(e). o0.8=α , before wandering correction (f).

o0.8=α , after wandering correction

Page 10: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(g).

o0.12=α , before wandering correction (h). o0.12=α , after wandering correction

Fig 7: Normarized T.K.E. distributions ( cx 0.4= , .1015.1Re 5×=C ).

Figure 8 shows the Reynolds Stress distributions of the wingtip vortex before and after wandering correction. The left column of the plots shows the distributions without wandering correction and the right column shows the distribution with wandering correction. Before correcting the center, the vortex core has two pairs of positive and negative Reynolds Stress except for at o0.2=α .These pairs of stresses become stronger as the angle of attack increases. For o0.8=α (Figure 8e), the maximum stresses are about ±2.0×10-4 and when the angle of attack is increased to o0.12=α , as shown in Figure 8g, the stress level exceeds ±3.0×10-4. These two pairs of positive and negative stresses are found in the study done by Chow et al. (1997). They describe that the contours of the Reynolds stress in the wake planes had a four-leaf clover pattern with alternately changing sign of stresses in each leaf. Each leaf was roughly aligned at ±45⁰ off the y and z axis, with positive levels of stress found in the first and third quadrants (+/+ and - values of z and y, respectively) and negative levels of stress found in the second and fourth quadrants. This four-leaf clover pattern is explained by converting the Reynolds stress equation from a Cartesian coordinate to the cylindrical coordinate system and the detailed explanation is described on Chow et al. (1997). The right column of Figure 8 shows the Reynolds stress distribution after the wandering correction. As shown in the figures, the Reynolds Stresses in the core region cannot be seen after the wandering correction. For the lower angles of attack, the stress level is near before the correction but after the wandering correction, the stress is completely neutralized to less than ±1.5×10-5 . For the higher angles of attack, especially at

o0.12=α , some stress is still remaining even after the wandering correction; however, the stress level is decreased more than 50%, from ±3.0×10-4 or greater to about ±1.5×10-4 . This reduction in the Reynolds stresses tells that the four-leaf clover pattern in the vortex core is mainly due to the vortex wandering. Interestingly enough, the wandering correction diminished the stress levels within the core and only within the core. For example, at o0.6=α , only the stress pairs within the core disappear and the stresses within the spiral wake, especially the square shaped positive stress region, stay almost the same in the terms of the magnitude and distribution pattern.

Page 11: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(a).

o0.2=α before wandering correction (b). o0.2=α , after wandering correction

(c).

o0.6=α before wandering correction (d). o0.6=α after wandering correction

(e).

o0.8=α before wandering correction (f). o0.8=α after wandering correction

Page 12: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(g).

o0.12=α before wandering correction (d). o0.12=α after wandering correction

Fig 8: Normarized Reynolds stress distributions ( cx 0.4= , .1015.1Re 5×=C ).

C. Comaprsions of the SPIV measurements and the predictions of the Devenport’s Method The Devenport’s method was applied to process the measurement date at two different angles

of attack, α= 2.0⁰ and 4.0⁰. Figure 9a shows the tangential velocity profiles of the vortex along the positive z-axis at α= 2.0⁰ and Figure 9b shows the tangential velocity profile with the same flow conditions at α= 4.0⁰. For both figures, the blue solid curve represents the velocity profile with vortex wandering. The black solid curve represents the velocity profile with wandering correction, i.e., instantaneously corrected wandering free velocity profile, and the red symbols represent the velocity profile estimated by Devenport’s method for wandering correction, i.e., numerically derived wandering free velocity profile. At α = 2.0⁰, as shown in Figure 9a, the SPIV method and Devenport’s method agree very well with each other from the vortex core center to the core radius. After passing the core radius, the Devenport’s correction curve starts to diverge from the SPIV curve and converging to the non-corrected curve. This is because the equations used to derive the Devenport’s curve are formulated to converge to the non re-centered profile at the outer region of the vortex. For this case, the SPIV corrected curve is more like the off–set profile from the non re-centered profile, so that there will be some discrepancy between those two curves at the outer region of the vortex. Figure 9b shows the velocity profile at α= 4.0⁰. For this case, the SPIV corrected profile and Devenport’s corrected profile within the vortex core do not match to each other, or the Deveport’s method recovered less tangential velocity than SPIV method. After passing the vortex radius, all three velocity profiles converged to each other indicating that there is not much wandering effect occurring at this particular outer region of the vortex core. By observing the profiles on Figures 9a and 9b, the Devenport’s method seems to be not consistently predicting the correct velocity profile in radial direction. For α= 2.0⁰ case, they have a good agreement within the core but for α= 4.0⁰ case they have a good agreement at the outer core region. This inconsistency happens because the Devenport method uses only one velocity profile on the positive z axis. This is not sufficient to fully recover the wandering free velocity profile accurately. One of the assumptions made by Devenport et al. (1996) was to assume that a wingtip vortex would be axisymmetric, however, in reality, a wing-tip vortex is

Page 13: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

not perfectly axisymmetric, as shown in Figure 10, and has the distinctive profile shape depending on the profile extracting locations.

0

0.2

0.4

0.6

0.8

1.0

0 1 2 3 4

SPIV re-centeredDevenport re-centeredNot re-centered

z/rc

v/v θ

0

0.25

0.50

0.75

1.00

0 1 2 3 4

SPIV re-centeredDevenport re-centeredNot re-centered

z/rc

v/v θ

(a). .deg0.2=α (b). .deg0.4=α

Fig. 9: Tangential velocity profile for non re-centered data and re-centered data using SPIV method and Devenport’s method ( 51015.1Re ×=C , 0.4=c ).

0

0.5

1.0

1.5

2.0

2.5

0 10 20 30

Positive y-axisPositive x-axisNegative y-axisNegative x-axis

r (mm)

v θ (

m/s

)

Fig. 10: Tangential velocity profiles along different axes (α=2.0⁰, Rec=1.15×105, x/c=4.0)

D. Effects of the Wind Tunnel Walls As described in the set-up section, an additional wall was inserted in the wind tunnel test

section to change the wall-to-tip distance from 6.0in to 4.0, 2.0, and 1.0in in order to investigate the effects of the wind tunnel walls on the evolution of the wingtip vortex. Figure 11 shows the projection of the vortex centers and the average vortex center with different wall lengths. When there is no additional wall, i.e., 6.0in wall-to-tip distance, the vortex center lies around the

Page 14: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

location of z /c= −0.052, which is near the trailing edge height. As the tunnel wall gets closer to the wing tip, the vortex starts moving upward. When the wall exists, it acts as the mirror vortex image which exists on the other side of the wall. Therefore, there will be a stronger influence of the vortex up-wash, which drives the vortex upward. The correlation between the wall to the wing-tip distance and the vortex up wash driven height can be shown in log-log scale plot as in Figure 12. By fitting the curve with the power equation, the equation is determined as ln y = −1.69 log x + 8.61. These coefficients may vary with the flow condition and wing angle of attack, as the equation is depending on the vortex radius.

+

y/c

z/c

-0.1 -0.05 0 0.05 0.1 0.15 0.2

-0.05

0

0.05

0.1

0.15

Tip to Wall = 1.5c

Tip to Wall = 0.25c

Tip to Wall = 0.50c

Tip to Wall = 1.0c

NACA0012Trailing Edgeat α=5.0°

Wind tunnel center

Fig. 11: Mean vortex location with different tip-to-wall distance.

( deg,0.5=α cx 0.4= )

Fig 12: Tip-to-Wall distance versus measn vortex center location in z-direction.

Z=0.0mm at airfoil trailing-edge. ( deg,0.5=α cx 0.4= )

Page 15: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

The Figure 13 shows the normalized T.K.E. and Reynolds stresses distributions of the wingtip vortex with the different wingtip-to-wall distances. As the wall gets closer from 4.0in to 1.0in, the T.K.E. values get higher. After removing the wandering effect, which is shown in Figures 13a and 13c, it is clearly shown that moving the inner wall closer to the wing-tip will increase the turbulence level within the vortex core more than twice. When the wall is further pushed up to the 1.0in tip-to-wall distance, the both non re-centered and re-centered distribution look the same. That suggests that the T.K.E., at 1.0in distance, is dominated by the turbulence within the vortex itself and the turbulence by the wandering has little or no effect impact on the total T.K.E. distributions. Table 2 shows the wandering statistics of the wing-tip vortex with the wall effect. The introduction of the additional wall increased the wandering amplitude about twice from 4.0≈σ to 7.0≈σ , and the center location distribution is less axisymmetric, e.g.,

cy r203.0=σ and cz r296.0=σ with 2.0in wall-to-tip distance. For the 1.0in tip-to-wall distance case, the vortex wanders 50% more in the y-axis direction. The vortex radius tends to get larger from 4.9mm to 6.33mm as the wall get closer from thirty to less than ten radii away from the tip; however, at 1.0in wall-to-tip, the radius gets smaller because the physical wall is getting closer to less than five vortex radii away. While the vortex behaves differently with the additional walls, the vorticity is not affected by the wall effect and stays around 1.3−≈ω at up to 8.0rc. When the wall is moved closer to less than 5 radii away from the tip, the vorticity gets smaller by 10%. Observing from the vortex structure and behavior, the vortex seems to interact with the mirror image vortex on the other side of the wall up to 8.0rc away; however, if the wall gets to within 5.0 or 4.0 radii way from the wing-tip, vortex would have interaction with the physical wall more and behaves more in a chaotic fashion.

Table2: Numerical data for SPIV method.

Page 16: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

(a). 4.0in tip-to-wall, T.K.E. (b). 4.0in tip-to-wall, Reynolds stress.

(c). 2.0in tip-to-wall, T.K.E. (d). 2.0in tip-to-wall, Reynolds stress.

(e). 1.0in tip-to-wall, T.K.E. (f). 1.0in tip-to-wall, Reynolds stress.

Fig 13: T.K.E. and Reynolds stress distributions with different tip-to-wall distance (α=2.0⁰, 51015.1Re ×=C , 0.4/ =cx ).

Page 17: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

Concluding Remarks In the present study, a high-resolution Stereoscopic Particle Image Velocimetry (SPIV)

system, which is capable of achieving quantitative measurements of all three components of instantaneous flow velocity vectors, was used to conduct detailed flow field measurements to quantify the transient behavior of the wing-tip vortex generated from a rectangular NACA0012 airfoil. One of the great advantages of SPIV over the classical technique is that the vortex wandering can be removed by tracking the center of the vortex in every SPIV frame. By tracking the center of the vortex, the effects of the wandering and turbulence on the evolution of the wingtip vortex can be separated. The results show that after wandering correction, the T.K.E. and Reynolds stress distributions become lower by more than 2 factor of two at 4.0 chord length downstream. This suggests that the wingtip vortex itself is laminar after the rollup and the higher turbulence intensity in the vortex core, reported in past studies, is mainly due to vortex wandering.

The Devenport’s method under and/or over predicts the tangential velocity depending on the radial distance from the center and the prediction is not consistently accurate. This is because the calculation is depending on only one tangential velocity profile. This is not sufficient to fully recover the accurate wandering free velocity profile. Therefore, the Devenport’s method should only be limited to approximate the velocity profile and with reasonably larger wandering amplitude, perhaps 5/cr>σ , for satisfactory results since the smaller wandering amplitude means less smearing of the velocity by the wandering. If the accurate wandering free velocity profile is necessary, the spatially-resolved instantaneous non-intrusive technique like PIV or SPIV is desirable.

When the wind tunnel wall gets closer to the wing-tip vortex, it will increase the T.K.E. levels. When the wall is set further apart, the T.K.E. comes from the wandering and the vortex turbulence itself, but when the wall gets closer enough to about 4.5 vortex core radius, then the T.K.E. is mainly coming from the turbulence level of the vortex itself and there is a little or no impact from the wandering. As the wall get closer to the wing tip, the vortex wanders more and the vortex radius gets bigger hence the image vortex on the other side of the wall effect is big. But as the wall gets even closer to within the four or five vortex radii away, the vortex starts to wander in z-direction more with the radius and vorticity becoming smaller. When the wall exists, it acts as the mirror vortex image exists on the other side of the wall, therefore there will be a stronger influence of the vortex up-wash, which drives the vortex upward. The correlation between the tip-to-wall distance and the vortex up-wash driven height is found to be fitted well by using a power equation.

Acknowledgments The authors also want to thank Mr. Bill Rickard of Iowa State University for his help in conducting the experiments. This work is sponsored by the AFOSR, grant FA9550-08-1-0294.

Page 18: The Effects of Wind Tunnel Walls on the Near-field Behavior of a Wingtip Vortexhuhui/paper/2011/AIAA-2011-1165... · 2010-12-22 · The Effects of Wind Tunnel Walls on the Near-field

Reference Bailey and Tavoularis (2008). Measurements of the velocity field of a wing-tip vortex, wandering in grid turbulence. J. Fluid Mech., 601, 281-315. Devenport, W. J., Rife, M. C., Liapis, S. I., and Follin, G. J. (1996). The structure and development of a wing-tip vortex. J. Fluid Mech., 312, 67–106.

Gerontakos and Lee (2006). Near-field tip vortex behind a swept wing model. Exp. Fluids, Vol. 4, No.1, 141–155. Heyes, A. L., Hones, R. F., and Smith, D. A. R. (2004). Wandering of wing-tip vortices. In Proc. 12th Intl Symp. On the Applications of Laser Techniques to Fluid Mechanics, Lisborn, Portugal.

Hu, H., Saga, T., Kobayashi, T., and Taniguchi, N. (2001). A Study on a Lobed Jet Mixing Flow by Using Stereoscopic PIV Technique. Phy. Fluid, Vol. 13, No.11, 3425–3441.

Prasad A. K., and Adrian R. J. (1993). Stereoscopic Particle Image Velocimetry Applied to Fluid Flows. velocimetry in liquid flows. Exp. Flows, 15, 49–60.

Soloff, S. M., Adrian, R. J., and Liu, Z-c. (1997). Distorsion compensation for generalized stereoscopic particle image velocimetry. Meas. Sci. Technol., 8, 1441–1454.


Recommended