Vous êtes sur la page 1sur 55

Interaction of a Dynamic Vortex Generator with a

Laminar Boundary Layer

By

Erica Jeannette Cruz

A Thesis Submitted to the Graduate


Faculty of Rensselaer Polytechnic Institute
in Partial Fulfillment of the
Requirements for the Degree of

MASTER OF SCIENCE

Major Subject: AERONAUTICAL ENGINEERING

Approved by the
Examining Committee:

__________________________________
Matt Oehlschlaeger
Committee Chair

__________________________________
Onkar Sahni
Member

__________________________________
Jason Hicken
Member

Rensselaer Polytechnic Institute


Troy, New York

December 2015
(For Graduation December 2015)
Pro Q ue st Num b e r: 10159646

All rig hts re se rve d

INFO RMATIO N TO ALL USERS


The q ua lity o f this re p ro d uc tio n is d e p e nd e nt up o n the q ua lity o f the c o p y sub m itte d .

In the unlike ly e ve nt tha t the a utho r d id no t se nd a c o m p le te m a nusc rip t


a nd the re a re m issing p a g e s, the se will b e no te d . Also , if m a te ria l ha d to b e re m o ve d ,
a no te will ind ic a te the d e le tio n.

Pro Q ue st 10159646

Pub lishe d b y Pro Q ue st LLC (2016). Co p yrig ht o f the Disse rta tio n is he ld b y the Autho r.

All rig hts re se rve d .


This wo rk is p ro te c te d a g a inst una utho rize d c o p ying und e r Title 17, Unite d Sta te s Co d e
Mic ro fo rm Ed itio n Pro Q ue st LLC.

Pro Q ue st LLC.
789 Ea st Eise nho we r Pa rkwa y
P.O . Bo x 1346
Ann Arb o r, MI 48106 - 1346
Copyright 2015

By

Erica Jeannette Cruz


All Rights Reserved

ii
CONTENTS

LIST OF FIGURES ............................................................................................... iv


ACKNOWLEDGMENTS .................................................................................... vii
ABSTRACT ......................................................................................................... viii
1. Introduction and Historical Review ...................................................................1
1.1 Passive Flow Control: Vortex Generators ................................................2
1.2 Active Flow Control: Vortex Generators .................................................3
2. Experimental Setup and Techniques ..................................................................4
2.1 RPI Subsonic Research Wind Tunnel ......................................................4
2.2 Dynamic Vortex Generator ......................................................................5
2.2.1 Actuator Design and Experimental Model ...................................5
2.2.2 Dynamic Vortex Generator Quantification ..................................6
2.3 Oil Flow Visualization .............................................................................8
2.4 Stereo Particle Image Velocimetry...........................................................8
2.4.1 Volume Reconstruction ..............................................................12
3. Results and Discussion ....................................................................................13
3.1 Oil Flow Visualization Qualitative Results .........................................13
3.2 SPIV Quantitative Results ...................................................................15
3.2.1 Time-Averaged Flow Field ........................................................15
3.2.2 Phase-Averaged Flow Field .......................................................29
3.2.3 Turbulent Kinetic Energy ...........................................................38
4. Summary and Conclusions ..............................................................................40
5. Future Work .....................................................................................................42
6. References ........................................................................................................43

iii
LIST OF FIGURES

Figure 2.1: CAD rendering of VG actuator and wind tunnel test section equipped
with VG plug and pressure containment box. .................................................. 5

Figure 2.2: Photo of wind tunnel test section equipped with the circular plug insert
and pressure containment box. ........................................................................ 6

Figure 2.3: Quantification of the DVG for a range of frequencies at various input
voltages. Data were normalized with respect to maximum voltage value. ..... 7

Figure 2.4: Schematic of the SPIV setup (top view, not drawn to scale).
Representative planes shown using the dashed lines. .................................... 10

Figure 2.5: Photo of experimental setup with the circular plug insert and vortex
generator. Ruler has been included to show scale. ........................................ 11

Figure 3.1: Oil flow visualization for static vortex generator as flow travels from
left to right. Major flow features identified. .................................................. 13

Figure 3.2: Iso-contours of normalized streamwise vorticity, , under static


(a) and dynamic (b) conditions at four streamwise locations, x/h = 8, 12, 16,
and 20 in 3D configuration. ........................................................................... 16

Figure 3.3: Wall-normal distributions of time-averaged streamwise velocity,


, at four streamwise locations, under static (a) and dynamic (b)
conditions, at the core of main vortex. .......................................................... 17

iv
Figure 3.4: Wall-normal distributions of time-averaged streamwise velocity,
, under static and dynamic conditions at four streamwise locations, x/h
= 8 (a), 12 (b), 16 (c) and 20 (d), at the core of main vortex. ....................... 18

Figure 3.5: Wall-normal distributions of time-averaged spanwise velocity, ,


under static and dynamic conditions at four streamwise locations, x/h = 8 (a),
12 (b), 16 (c) and 20 (d), at the core of main vortex...................................... 20

Figure 3.6: Iso-contours of normalized streamwise vorticity, , under static


(a-d) and dynamic (e-h) conditions at four streamwise locations, x/h = 8, 12,
16, and 20....................................................................................................... 22

Figure 3.7: Schematic of an array of co-rotating vortices being represented by four


potential vortices. ........................................................................................... 25

Figure 3.8: Iso-surfaces of Q-criterion colored by postive (red) and negative (blue)
time-averaged normalized streamwise vorticity, , for static (a) and
dynamic (b) cases. ......................................................................................... 27

Figure 3.9: Iso-contours of Q-criterion under static (a-d) and dynamic (e-h)
conditions at four streamwise locations, x/h = 8, 12, 16, and 20. .................. 28

Figure 3.10: Iso-contours of normalized streamwise velocity, , for time-


averaged static and dynamic cases (a) and phase-averaged dynamic case at
phase angles  = 0 (b), 60 (c), 120 (d), 180 (e), 240 (f), and 300 (g), at
streamwise location x/h = 16. ........................................................................ 30

Figure 3.11: Wall-normal distribution of normalized phase-averaged spanwise


velocity, , under dynamic conditions at single streamwise location (x/h
= 16), following core of main vortex, for six phases:  = 0, 60, 120, 180,
240, and 300. .............................................................................................. 32
v
Figure 3.12: Wall-normal distribution of normalized phase-averaged streamwise
velocity,  , under dynamic conditions at single streamwise location (x/h
= 16), following core of main vortex, for six phases:  = 0, 60, 120, 180,
240, and 300. .............................................................................................. 32

Figure 3.13: Iso-contours of normalized streamwise vorticity,  , for time-


averaged static and dynamic cases (a) and phase-averaged dynamic case at
phases  = 0 (b), 60 (c), 120 (d), 180 (e), 240 (f), and 300 (g), at
streamwise location x/h = 16. ........................................................................ 34

Figure 3.14: Iso-surfaces of Q-criterion colored by postive (red) and negative (blue)
normalized streamwise vorticity,  , for time-averaged static and
dynamic cases (a) and phase-averaged dynamic case at phases  = 0 (b), 90
(c), 180 (d) and 270 (e). .............................................................................. 37

Figure 3.15: Wall-normal distributions of total kinetic energy, TKE, under static
and dynamic conditions at four streamwise locations, x/h = 8 (a), 12 (b), 16
(c) and 20 (d), at the core of main vortex. ..................................................... 39

vi
ACKNOWLEDGMENTS

This work has been supported by the Boeing Company, monitored by Dan Clingman. I
would like to thank Matt Oehlschlaeger for chairing this committee. I would like to
acknowledge my advisor, Prof. Michael Amitay for giving me the opportunity to conduct
this research, as well as Drs. Burak Tuna, Keith Taylor, and Chia Leong for their support.
I would like to thank my fellow graduate students for many lessons learned in a research
environment. Furthermore, undergraduate student Mr. Wilfred Chan must be recognized
for his tremendous help with the design, building, and testing of the actuator used in this
research.

vii
ABSTRACT

An experimental investigation was performed to study the fundamental interaction


between a static and dynamic vortex generator with a laminar boundary layer. The
effectiveness of static vortex generators (VGs) on delaying boundary layer separation is
well established. However, as a passive flow control device, static VGs are associated with
a drag penalty since they are always present in the flow. In the current study a piezoelectric-
based dynamic vortex generator (DVG) was developed with the goal of mitigating the drag
experienced when using a VG as a flow control device and exploring whether or not a DVG
was more effective in flow mixing within the boundary layer. Experiments were conducted
in a small wind tunnel, where the VG was flush mounted to the floor. The VG was
rectangular in shape and erected into the flow with a mean height of the local boundary
 
   = 3 mm. The skew angle of the VG was   18 with respect to
the incoming flow, oscillated at a driving frequency of  = 40 Hz with a peak to peak
displacement (or amplitude) of    = 1.5 mm. During the experiments, the free
stream velocity was held constant at  = 10 m/s. This corresponded to a Reynolds number
of   2000, which was based on the local boundary layer thickness at the center of the
VG. Surface oil flow visualization experiments were performed to obtain qualitative
information on the structures present in the flow, while Stereoscopic particle image
velocimetry (SPIV) was used to provide quantitative measurements of the 3-D flow field
at multiple spanwise planes downstream of the VG under both static and dynamic
conditions. Several flow features were detected in the oil flow visualization experiments,
including two vortical structuresthe main vortex and primary horseshoe vortexwhich
were confirmed in the SPIV results. The time-averaged flow field showed similar results,
though the strength of the vortices appeared less when the VG was actuated. However,
phase-averaged data revealed the size, strength, and location of the vortices varied as a
function of the actuation cycle, with peaks of vorticity magnitude being greater at certain
phases as compared to the static case. The varying flow field associated with the dynamic
motion of the DVG showed higher levels of turbulent kinetic energy, therefore confirming
enhanced mixing in contrast to the static case.

viii
1. Introduction and Historical Review

Improving aerodynamic performance can be achieved in two ways the design of


the vehicle can be optimized or the flow moving around the vehicle can be controlled. The
latter is achieved through the use of either passive or active flow control devices. A vortex
generator, VG, is a passive flow control device that has been used extensively on wings to
control flow separation through the introduction of a streamwise, or longitudinal, vortex.
This vortex is formed by placing a VG on a surface with a height such that it is embedded
within the boundary layer and a skew angle with respect to the incoming flow. Due to the
pressure differences between the two sides of the VG and its finite height, the incoming
boundary layer rolls up into a single streamwise vortex downstream. This allows high
momentum fluid in the outer flow to be drawn into the lower momentum boundary layer.
This transfer of momentum energizes the boundary layer, therefore helping to delay or
completely mitigate flow separation. This method of flow control has proven to be effective
in the presence of flow separation (e.g. at high angles of attack or over control surfaces).
However, due to the nature of a passive flow control device, as it is always present in the
flow regardless of operating conditions, there is an associated drag penalty. This was the
motivation to explore the feasibility and effectiveness of deployable or dynamic VGs, or
DVGs. There are two main features associated with the DVGs. First is, they can be deployed
as needed and retracted when not. Secondly, these DVGs can be oscillated from specified
offsets at different frequencies. The intent of the current study to explore the fundamental
interaction of a single DVG with a cross flow over a flat plate. With a more thorough
understanding of the flow physics associated with the dynamic motion of a DVG, it is
hypothesized that they have a strong potential for use as active flow control devices on
aircraft. It is predicted that the performance of a VG can be enhanced through increasing
the flow mixing near the surface. This can be maximized by actuating the DVG at certain
frequencies associated with the flow. This capability can then be used to help increase the
overall aerodynamic efficiency of the aircraft, with reduction in drag and fuel consumption.
Ultimately, DVGs can be used to help further delay flow separation, leading to decreased
takeoff and landing distances, and    
  

1
1.1 Passive Flow Control: Vortex Generators

The use of conventional passive VGs was first introduced by H.D. Taylor at United
Aircraft in 1947 to prevent boundary layer separation in a wind tunnel diffuser. This led to
the first systematic study of VGs and their effect on the boundary layer conducted by
Schubauer and Spangenberg in 1959 (10). Since then, VGs have been the subject of
extensive experimental and numerical investigations in both internal and external flows.
Internal flows refer to inlet ducts and diffusers, in which case VGs are used to prevent flow
separation and reduce total pressure distortion. In most cases, VGs are used in external
flows, such as on airfoils. In this case, they are used to delay flow separation over a surface
or enhance aircraft lift. This is done by controlling the boundary layer by redirecting the
outer flow towards the wall using the streamwise vortices created by vortex generators
(Schubauer, 12).

Though VGs have been found to be efficient in the transportation of high


momentum fluid of the free stream towards the lower momentum region near the wall, they
are associated with a large amount of parasitic drag as they are always present in the flow.
In an effort to reduce this drag, while still improving aerodynamic efficiency was to
generate an embedded streamwise vortex using a VG that was significantly smaller in
height. In the late 1980s, Rao and Kariya focused on VGs that were submerged within the
boundary layer (823). Experiments were conducted with a separated flat plate boundary
layer and different configurations were tested with a VG height of the order of 60% of .
The conclusion was that these smaller VGs were most effective in a counter-rotating
configuration and smaller sizes should still be investigated. Lin et al. showed that a VG
with the height of 10 to 20 percent of the boundary layer thickness was still effective in
transporting sufficient momentum from the outer into the boundary layer in order prevent
boundary layer separation (Lin, 4). Sub-boundary layer VGs have also been shown to
substantially increase the lift-to-drag ratio on multi-element high-lift airfoils (Lin, 1317) or
to effectively control separation in an adverse pressure gradient (Jenkins, 3).

2
1.2 Active Flow Control: Vortex Generators

An alternative method to mitigating the drag penalty associated with VGs is through
the use of deployable or dynamic VGs, or DVGs. In the case of deployable VGs, the device
would be deployed as needed (e.g. during takeoff and landing), but then be retracted during
cruise conditions when flow control is not necessary.

In addition to deploying the VG as needed, it is hypothesized that the effectiveness of


the VG can be enhanced through oscillations at frequencies present in the flow.
Dynamically driven vane-type VGs have recently been investigated by Osborn et al. on the
suction surface of a deflected trailing-edge flap. The DVG in this experiment could reach
a maximum frequency of 70 Hz when driven by an appropriately sized voice-coil actuator.
They proved that an array of DVGs was effective in mitigating flow separation, whereas a
similar array of static VGs was not. Subsequently, Seshagiri et al. investigated the effects
of a static vs. a dynamic VG on an airfoil at a low Reynolds number to increase the
maximum lift coefficient and the stall angle. Interestingly, they showed that DVGs in their
present configuration, did not appear to be effective compared to the static VGs.

In more recent years, Barth et al. studied the effect of vortices generated by a static
versus a dynamic VG in a turbulent boundary layer. They discovered that when the DVG
oscillated between the surface plane and a height of 2 mm above it, the vortices decayed
faster than the ones generated by the static VG. Le Pape et al., on the other hand, presented
a study on using leading edge DVGs for dynamic stall control. Their results showed that
different compromises between lift and pitching moments could be achieved depending on
the phase actuation of the DVGs.

3
2. Experimental Setup and Techniques

2.1 RPI Subsonic Research Wind Tunnel

Experiments were performed in an open return, small scale, low turbulence, low
speed, suction-based wind tunnel facility at the Center for Flow Physics and Control
(CeFPAC)     
              9:1
with a length to diameter ratio of 1.5. The cross-section is 0.1 m by 0.1 m with a test section
length of 0.61 m. Located upstream of the test section area are screens and a honeycomb
which condition the flow ahead of the test section, yielding turbulence intensity values
below 0.5%. The flow was driven by an AC motor, which was connected to a variable
frequency drive, providing a maximum free stream velocity of ~40 m/s. All experiments
presented herein were conducted with a constant free stream velocity of  = 10 m/s,
corresponding to a local boundary layers thickness at the center of the VG of about   3
mm. The resulting Reynolds number based on boundary layer thickness,  , was about
2000. Using stability analysis, the frequency of the most amplified Tollmien-Schlichting
(T-S) wave at the location of the DVG was determined to be approximately 170 Hz with
unstable frequencies ranging from 140 to 200 Hz.

Three of the four walls of the test section (except for the floor), were constructed of
acrylic glass allowing for the optical access required for the oil flow visualization and
stereoscopic particle image velocimetry, SPIV, experiments performed in this study. The
floor of the test section was equipped with a single row of pressure ports at the mid-span
of the floor to monitor the pressure gradient. The floor also had a circular cutout at the
upstream portion of the test section, which allowed the VG model to be flush-mounted to
the test section through a circular plug insert.

4
2.3 Oil Flow Visualization

A preliminary qualitative experiment was performed on the floor of the wind tunnel
using oil flow visualization to identify the location of flow structures. This technique
indicates the separation lines as the oil cannot penetrate into the separated region. The oil
used in this experiment was a mixture of heavy viscosity silicone oil and a fluorescent
tracer, which is best illuminated in the green wavelength under the excitation of a UV light.
The oil was distributed evenly both upstream and downstream of the static VG before
beginning the experiment. Images were taken using a digital SLR camera (Nikon D40)
with a long exposure time. A 60 mm focal length lens was used along with a 532 nm +/-
10 nm bandpass filter for image acquisition. A series of images was acquired every one
minute to observe the progression of the oil until a steady state was reached after about 20
minutes. Another purpose for utilizing this technique was to help determine the
measurement domain for the SPIV experiments.

2.4 Stereo Particle Image Velocimetry

After conducting the oil flow visualization experiments, the SPIV technique was used
to acquire spatially-resolved quantitative flow patterns. The SPIV system was a commercial
LaVision system. Major components of this system include a dual pulsed Nd:YAG laser
with a maximum output of 120mJ/pulse, two 1376 x 1040 pixel resolution thermos-
electrically cooled 12-bit Imager Intenser CCD cameras, and a programmable timing unit.
The cameras were mounted on an optical table. A -50 mm cylindrical lens was used to
create the light sheet in conjunction with a focal lens to focus the sheet at the measurement
domain to a thickness of about 1 mm at its waist. The laser sheet was positioned
perpendicular to the flow direction and permitted to shine through the acrylic glass
sidewalls of the wind tunnel test section. Using a computer-controlled three-axis traversing
system mounted on an optical table and below the test section, the laser sheet was aligned
with the area of interest, based on results from the oil flow visualization experiment. The
flow was seeded with O( ) water-based smoke particles, generated by a Martin
Magnum 800 theatrical fog machine. The smoke was injected into the tunnel upstream of

8
the test section to allow the particles to fully mix into the flow for more complete seeding
and become more uniformly distributed at the location of the laser sheet.

For the wind tunnel SPIV, the cameras and the laser optical head were each rigidly
attached to a three axis Velmex Bi-Slide stepper motor controlled traversing system, which
has a positional accuracy of 4m. This positioning system allowed for the cameras and
the laser to be uniformly translated after the system was focused and calibrated, which
enabled multiple spanwise data planes to be collected at incremental streamwise positions
so that a full three dimensional time-averaged fluid volume could be reconstructed. The
data collection and reconstruction method is discussed in more detail in the following sub-
sections.

The schematic of the SPIV setup along with its coordinate axes orientation can be
seen in Fig. 2.4. The origin of the measurement domain is located at the center of the VG,
where the x-axis is in the streamwise direction, the y-axis is normal to the tunnel floor and
the z-axis is in the spanwise direction. SPIV images were acquired at 25 spanwise planes,
which ranged from 20 mm downstream of the DVG, i.e., x = 20 mm, to 70 mm downstream
of the DVG   
                 

mm), 2 mm increments in the mid       50 mm), and 5 mm increments in the
far field of the DVG      

The cameras were positioned to capture the same field of view with an angular
separation between their viewing angles of 80 as shown in Fig. 2.4. The focal issues
associated with the angular misalignment were then accounted for by using a Scheimpflug
adapter between each camera and its lens as prescribed by Prasad and Jensen (3). A pair of
150 mm focal length lenses were used in conjunction with 532 nm 10 nm band-pass
filters on the cameras in order to acquire data in the spanwise oriented planes downstream
of the VG, with an interrogation window of about 25 mm x 25 mm. The cameras and laser
were controlled and synchronized using a LaVision PTU8 programmable timing unit.

9
2.4.1 Volume Reconstruction

Data on 25 spanwise planes were acquired along the center of the test section of the
wind tunnel, as previously described. Stereoscopic PIV data collected on these planes were
within a window size of about 25 mm x 25 mm. The data from all the planes were then
reconstructed into a volume to provide the 3-D interaction domain between the static and
dynamic vortex generator and laminar boundary layer.

12
From these results, several flow features can be observed. First, around the leading
edge and suction side of the VG, a small region of oil accumulation is present. This is
indicative of a leading edge separation bubble. Next, a separation line slightly below the
pressure side of the VG can be observed. It extends around the VG on the pressure side,
continuing downstream, and partially extends on the suction side. This separation line
represents the vorticity of the incoming boundary layer, which has evolved into a horseshoe
vortex. This is labeled as the primary horseshoe vortex. The incoming boundary layer,
upstream of the VG, is associated with spanwise vorticity, which wraps around the VG and
develops into a streamwise (or longitudinal) vortex downstream of the VG. Typically, an
incoming boundary layer upstream of an obstruction in the flow would develop into a pair
of counter-rotating vortices, which is characteristic of a horseshoe vortex. However, as
indicated by the presence of a leading edge separation bubble, the flow has separated at the
leading edge of the VG, hindering the formation of the other vortex on the suction side of
the VG. There is another separation line present that is slightly upstream of the VG,
extending around the pressure side of the VG and the primary horseshoe vortex. It also
partially extends around the suction side of the VG. This is a secondary horseshoe vortex,
created by the aforementioned process. Due to its distance away from the separated region
near the leading edge of the VG, this secondary horseshoe vortex is permitted to initially
form as is characteristic of a horseshoe vortex pair. However, on approach of the separation
flow region, the nearby vortex is no longer present. This is due to the pressure gradient in
this area which forces the flow to move towards the suction side of the VG where the flow
as separated and hence dissipates as did the primary horseshoe vortex. The strength of the
secondary horseshoe vortex has previously been found to be less coherent and in this case,
it seems to dissipate within two lengths of downstream of the VG. Due to the presence of
the main and primary horseshoe vortices, there is a region of induced velocity that is present
on the pressure side of the VG. As will be observed in subsequent sections, this region is
not indicative of another vortex.

Studying these results father downstream of the VG, it can be observed that the main
vortex separation line gets thicker. This suggests that the vortex created by the VG is a
coherent structure, but its strength dissipates as a function of downstream distance. The
primary horseshoe vortex also appears close to the main vortex at distances farther

14
downstream of the VG. The flow structures observed in the oil flow visualization
experiments were further explored using SPIV and will be discussed below.

3.2 SPIV Quantitative Results

A fundamental understanding of the interaction between a vortex generator, static or


dynamic, with a laminar boundary layer, requires the study of the three-dimensional flow
structures along a flat plate. Hence, the oil flow visualization results were coupled with
SPIV measurements to obtain temporally- and spatially-resolved quantitative data. The
qualitative results from the oil flow visualization experiment provided a guide for the
measurement/interrogation domain for SPIV measurements. Results collected from 25
spanwise planes were reconstructed into two-dimensional flow fields and three-
dimensional flow volumes to demonstrate both the complexity and three-dimensionality of
the flow field. The cases described herein are with the VG at a skew angle of  = 18 , a
mean height of  = 3 mm, a driving frequency of  = 40 Hz, and a peak-to-peak
displacement (or amplitude) of  = 1.5 mm.

3.2.1 Time-Averaged Flow Field

A quantitative comparison of the time-averaged normalized streamwise velocity


field,  , between the static and dynamic cases are shown in Fig. 3.2 at four streamwise
locations (i.e. x/h = 8, 12, 16, and 20) downstream of the VG. Results from the static and
dynamic cases are presented in Fig. 3.2a and b. respectively. In these figures, contours
denote normalized streamwise velocity. Lighter and darker colors correspond to higher and
lower velocity values, respectively. The free-stream velocity,  , is along the x-direction
and is moving from the lower left to the upper right. The axes have been normalized with
respect to the mean height of the VG,  = 3 mm. From these contour plots, the presence
of a velocity deficit core can be observed, which is consistent with the presence of the
vortexexpected to be formed in the wake of a vortex generator. In both cases, the main
vortex grows in size as a function of downstream distance. Moreover, with increasing
downstream distance, the main vortex convects in the spanwise direction (i.e. z-direction).

15
layer. The increased fluid mixing that is occurring due to the transport of high momentum
fluid from the free stream into the lower momentum fluid near the wall is causing a
transition of the laminar boundary layer into the turbulent regime. Further evidence of this
transition is the decrease in velocity gradient observed near the wall.

Supplementary to the streamwise velocity profiles seen in Fig. 3.4, Fig 3.5 provides
the wall-normal distributions of the time-averaged normalized spanwise velocity,  ,
at the same streamwise locations. Presented in these plots are three cases: baseline (i.e. no
VG is present), time-averaged static case, and the time-averaged dynamic case. When there
is no VG present in the flow, the results show an average spanwise velocity of zero,
demonstrating no anomalies in the flow field. When a VG is present in the flow, either
static or dynamic, there are both positive and negative spanwise velocities. The greatest
magnitude of positive spanwise velocity occurs in the near field of the VG, decreasing with
downstream distance, whereas the opposite occurs for negative spanwise velocity. The
point at which the spanwise velocity converts from negative to positive is the location of
the vortex core. This location increases in the wall-normal direction as a function of
downstream distance, showing that the main vortex core is moving upward, away from the
tunnel floor. Immediately downstream of the VG, the resulting vortex is embedded within
the boundary layer as seen in Fig. 3.5a. As the vortex develops farther downstream and
grows in size, higher momentum fluid is being moved from the outer flow towards the
wall. While the vortex core is still within the viscous region, the upper portion of the vortex
is still able to transfer more momentum from the outer flow. This creates higher magnitudes
of positive spanwise velocity in this area as compared to the lower portion, where the no-
slip condition at the wall reduces the fluid velocity. However, as is observed in Figs. 3.5b
 d, the magnitude of the negative spanwise velocity is increasing, due to this transfer of
momentum into the boundary layer. Moreover, there is an interaction between the main
and primary horseshoe vortices, which behave as a pair of co-rotating vortices. The primary
horseshoe vortex is smaller in size compared to the main vortex. As fluid moves around
the smaller vortex, it entrains the up-wash fluid moving from under the main vortex. As
will be presented in forthcoming vorticity contour plots, there is an exchange of vorticity
that is the basis of this interaction, helping to hasten the negative spanwise velocity near
the wall surface.

19
the smaller horseshoe vortex when the VG is actuated, resulting in less fluid being drawn
in the negative spanwise direction. This suggests unsteadiness in the formation of the
horseshoe vortex which will be furthered explored in the discussion of the phase-averaged
results. When studying the results further downstream, the negative spanwise velocity near
the wall are similar whether or not the VG is being actuated. This means the effects of the
dynamic motion of the VG on the fluid near the wall surface are limited to immediately
downstream of the actuated VG.

Velocity fields are helpful in understanding flow fields, however, in the case of
identifying possible vortical structures, examination of vorticity contour plots are required.
Figs. 3.6a-d and 3.6e-h present the time-averaged normalized streamwise vorticity,
  , contour plots for the static and dynamic cases, respectively. These plots represent
normalized streamwise vorticity fields in the same streamwise locations as previously
presented (i.e. x/h = 8, 12, 16, and 20) and the flow is going into the page as one is looking
from upstream to downstream. Solid and dashed lines represent positive (clockwise) and
negative (counterclockwise) vorticity, respectively, and darker colors indicate larger
magnitudes of vorticity. Most notable in the time-averaged vorticity contour plots is the
large presence of strong positive vorticity near z/h = -1.0. This is the main vortex observed
in the plots of velocity deficit in the streamwise contours previously presented. This vortex
is initially embedded within the boundary layer, where vorticity is rapidly diffused and the
vortex grows. As mixing increases and the boundary layer transitions from laminar to
turbulent, the vortex is further distorted. As described by Perkins, this diffusion of vorticity

    ) and in the
is caused by strong gradients in the normal stress anisotropy (
   ). The gradients in these quantities appear in the mean
crossflow-plane shear stress (
transport equation for streamwise vorticity, and contribute to the decay of streamwise
vorticity as described by Perkins (724). There is a loss of energy that goes into creating and
sustaining the vortex while it develops farther downstream.

In addition to the large region of positive vorticity observed in both the static and
dynamic cases, another smaller concentration of positive vorticity in the near-field (i.e. x/h
= 8 and 12) can be seen alongside the main vortex in the spanwise direction. This is the
primary horseshoe vortex, which developed as the incoming boundary rolled up on
approach of the VG. This creates a pair of co-rotating vortices in the flow field, which has

21
measurement domain. A review of the results from the oil flow visualization experiment
reveals that the horseshoe vortex does indeed extend farther along the test section. In the
dynamic case, the vorticity concentration of the horseshoe vortex is comparable to static
conditions, however, there is minimal interaction between the two vortices. Furthermore,
the spanwise distance between the two vortices is greater in the dynamic case as compared
to the static counterpart. Once the vortices have traveled to the mid-field region in the
dynamic case, the horseshoe vortex has moved out of the measurement domain. It can be
hypothesized that this vortex will continue to develop farther downstream, though its
vorticity will continue to dissipate and have little to no interaction with the main vortex.

Next, there is a region of negative vorticity near the wall which grows in magnitude
with downstream distance. The interaction of the spanwise and normal velocity
components, and the no-slip condition at the wall produces the region of negative vorticity
below the main vortex. As high momentum fluid is being transported from the outer flow
to the wall, the flow field is being altered by increasing the velocity gradients experienced
in this area. This negative vorticity is convected by the main vortex and accumulates on
the up-wash side of the vortex. As the main vortex develops farther downstream, the
accumulation of negative vorticity near the wall grows. This region of negative vorticity
near the wall does not represent another vortex, but only velocity gradients caused by the
aforementioned interaction. This accumulation of opposite vorticity to those of the main
and horseshoe vortices contribute to the enhance mixing that is occurring within the
boundary layer. As was observed in the streamwise velocity profiles, this results in a
transition from a laminar to a turbulent boundary layer. Furthermore, the negative vorticity
causes the main vortex to move upwards, away from the wall, which was detected in the
spanwise velocity profile plots. Comparison between the time-averaged static and dynamic
cases show that the accumulation of negative vorticity near the wall is greater when the VG
is not actuated. This is due to the lesser magnitude of vorticity of the main vortex in the
dynamic case in a time-averaged sense a direct result of increased exchange of
momentum in the dynamic case, as previously discussed and will be further explored in
subsequent sections while examining the phase-averaged results. As a second order effect
of the main vortex having less vorticity and therefore resulting in less accumulation of
negative vorticity is that the interaction between the main and horseshoe vortices is
decreased. As the negative vorticity near the wall increases, the rotating flow about the

23
horseshoe vortex is entrained by the increased velocity gradients, drawing the horseshoe
vortex closer to the main vortex. This is clearly observed by the differences is spanwise
distance between the two vortices when comparing the static and dynamic cases. This
relationship will be more precisely defined when investigating the phase-averaged
streamwise vorticity contour plots.

In addition to the aforementioned areas of negative vorticity, there are small


concentrations located above the vortices. Similar to the negative vorticity concentrations
near the wall, there is also an interaction between the outer flow and the vortices, creating
velocity gradients which correspond to opposite vorticity. These concentrations, however,
are only present in the static case. Due to the increased dissipation of vorticity in the
dynamic case resulting from enhanced momentum exchange, the flow field in the areas
above the vortices are unable to interact strongly enough such that the nearby velocity
gradients create sufficient concentrations of opposite vorticity.

The regions of opposite vorticity near the main vortex also contribute to the
stretching of the main vortex. The divergent velocity field caused by these areas of opposite
vorticity act to decrease the streamwise component of vorticity. This is a result of the
conservation of angular momentum. As the area of the vortex is increased, the vorticity
must decrease in order to keep the circulation constant. For this reason, more stretching of
the main vortex is observed in the time-averaged static case as compared to the dynamic
case. With increasing downstream distance, the concentration of negative vorticity above
the main vortex decreases, resulting in less vortex stretching.

Further observations from the streamwise vorticity contour plots in Figs. 3.6a-b
include the spanwise convection of both vortices. This movement can be explained using
potential flow theory. This array of co-rotating vortices can be represented by potential
vortices, as depicted in Fig. 3.7. This schematic represents the two vortical structures seen
in the vorticity contour plots. These vortices can then be represented with four potential
vortices, such that the no-penetration condition at the wall is satisfied. This results in
induced velocity in the spanwise direction. The induced velocity,  , at a given point is
affected by the strength of the vortex, , and its distance from the vortex, , hence affecting
the amount of spanwise convection of the vortices (equation 1). As a result, a co-rotating

24
The Q-criterion represents the local balance between strain rate, S, and vorticity
magnitude, . Hunt et al. identified vortices of an incompressible flow as fluid regions
where the flow field is dominated by rotation (Q > 0), making this method a strong indicator
of vortices and their strength. However, the Q-criterion only illustrates the presence of
rotation in a flow field as it is unable to indicate direction of the rotation. Therefore, to
determine the direction of fluid rotation, the Q-criterion is colored by the time-averaged
normalized streamwise vorticity,   . Presented in Figs. 3.8a-b are the time-averaged
iso-surfaces of Q-criterion for values 0.04 < Q < 0.10. Red and blue colors indicate fluid
rotation in positive and negative streamwise direction, respectively. Seen in this figure are
the presence of two developed vortical structures with positive rotation, under both static
and dynamic conditions. The larger vortex is the main vortex, whereas the smaller vortex
is the primary horseshoe vortex. In both cases, the size and relative location of the main
vortex are similar. As was observed in the vorticity contour plots, the horseshoe vortex is
not as coherent in the dynamic case. Under both static and dynamic conditions, there are
scattered regions of negative rotation, due to the interaction between the vortices and the
wall. In the static case, there are slightly more areas of organized negative rotation between
the main and horseshoe vortices. This corresponds to the accumulated negative vorticity
on the up-wash side of the main vortex, resulting in induced velocity in this area. This
accumulation was not sufficient to develop a fully induced vortex with rotation in the
opposite direction.

In order to gain a more in-depth look at the streamwise development of the vortices,
Figs. 3.9a-h present the iso-contours of the Q-criterion at the streamwise locations x/h = 8,
12, 16, and 20. In an effort to have direct comparisons between the different streamwise
locations, the same contour levels were used for each location. However, due to the rapid
dissipation of vorticity of these vortices in the streamwise direction, the number of contour
levels also quickly decrease. For this reason, the near-field contour plots have significantly
more levels of Q-criterion, making them slightly challenging to distinguish. Nevertheless,

26
3.2.2 Phase-Averaged Flow Field

An examination of the time-averaged flow field under static and dynamic conditions
has revealed various flow features and differences between the two cases, however the
picture thus far has been incomplete. The time-averaged results only depict the mean flow
field. Due to the sinusoidal actuation of the VG, it is necessary to explore the flow fields
when it is locked to this motion. This will allow for a deeper understanding of the effects
of the dynamic motion of the VG as the evolution of any coherent flow structures can be
tracked over the actuation cycle.

A comparison between the iso-contours of time- and phase-averaged streamwise


velocity,  , at the streamwise location of x/h = 16 is presented in Figs. 3.10a-g. Fig.
3.10a depicts the time-averaged results under both static and dynamic conditions. Rows of
images in Figs. 3.10b-g correspond to sequential phases during oscillation of the VG when
it is being actuated, designated as  = 0, 60, 120, 180, 240, and 300. The flow field
varies widely throughout the actuation cycle. In Fig. 3.10a, at the beginning, there is only
a slight presence of the main vortex and no velocity deficit core. As the actuation cycle
progresses, the distortion of the boundary layer becomes more pronounced. The velocity
deficit core is more prevalent in the later phases of the cycle, albeit, its strength and relative
location vary. Moreover, areas of accelerated flow can be tracked throughout the actuation
cycle. In Fig. 3.10c, accelerated flow can be seen on the downwash side of the vortex. That
flow is then observed under the main vortex (near the wall) in Fig. 3.10d. Later in the cycle,
the accelerated flow is moving away from the main vortex in a negative spanwise direction.
Finally, at  = 300, a distinct velocity deficit core remains. The phase-locked results reveal
how susceptible the flow field is to the dynamic motion of the VG. In the time-averaged
case, there is no distinct velocity deficit core at this particular streamwise location, making
it less evident that higher momentum fluid was being drawn to the boundary layer at the
magnitudes seen in the phase-averaged results. Furthermore, the varying location of the
accelerated flow during the actuation cycle suggest flow mixing is being extended over a
greater area under dynamic conditions. As was observed in the mean flow-fields, this
would result in less concentrations of fluid in the dynamic case in comparison to the static
case.

29
Comparable to the time-averaged streamwise data, in order to gain a more
quantitative understanding of the phase-averaged flow field, it is necessary to study the
velocity profiles. Shown in Fig. 3.11 is the wall-normal distribution of normalized phase-
averaged streamwise velocity,  , for six phases along the actuation cycle:  = 0, 60,
120, 180, 240, and 300. At the beginning and end of the cycle (i.e.  = 0 and 300),
the VG is fully embedded within the local bounday layer, at a height of approximately h
1.5 mm ( 
). Despite the small height, there is a noticeable impact on the flow field,
as a velocity deficit can still be observed, which extends towards the wall. At the middle
of the acutation cycle (i.e.  = 120 and 180), the velocity deficit peaks, along with a peak
acceleration near the wall. This corresponds to the region of accelerated flow seen in the
streamwise contour plots and is nearly a 20% increase in streamwise velocity as compared
to the beginning and ending phases.

Supplementary to the streamwise velocity profile plots are the wall-normal


distributions of normalized phase-average spanwise velocity,  , for the same phases,
which are presented in Fig. 3.12. Similar to the streamwise velocity profile for various
phases along the actuation cycle, the greatest magnitudes of spanwise velocity are observed
during the middle phases. From the changing locations and magnitudes in peaks of both
positive and negative spanwise velocity throughout the different phases, it is implied that
that the strength and location of the main vortex is also changing. As will be shown in
forthcoming vorticity contour plots, the negative spanwise velocity increase is associated
with the accumulation of negative vorticity near the wall. Resulting from the oscillating
magnitudes of spanwise velocity near the wall would be an increase in momentum transfer
and vorticity diffusion both of which are needed for enhanced flow mixing. This would
explain why the time-averaged flow fields depict lesser values for the same given
streamwise location as more energy has been used for the increased flow mixing.

31
For deeper insight into the flow physics associated with the vortical structures
present in the flow field, a comparison between the iso-contours of time- and phase-
averaged streamwise vorticity,   , at the same streamwise location (i.e. x/h = 16) is
presented in Figs 3.13a-g. Similar to Fig. 3.10, the time-averaged results (under both static
and dynamic conditions) and phase-averaged results are shown in Figs. 3.13a and 3.13b-g,
respectively. First, it can be observed from the phase-locked flow fields, that the vorticity
concentration of the main vortex varies as a function of the input phase angle, reaching its
peak vorticity near the middle of the actuation cycle (i.e.  = 180). Corresponding with
increased positive vorticity concentration of the main vortex, there is an increase in the
concentration of negative vorticity near the wall. As described when comparing the time-
averaged static and dynamic cases, increased vorticity concentration of the main vortex
and the near wall region has an effect on spanwise distance between the main and horseshoe
vortices. From the phase-averaged results, the evolution of the vortices can be better
explained. First, at the beginning of the actuation cycle, both the main and horseshoe
vortices are present, thought the strength of the latter is minimal. With the upstroke of VG,
the size and strength of the main vortex increases, forcing the horseshoe vortex away (and
outside the measurement domain). However, with the strength increase of the main vortex,
there is an associated accumulation of negative vorticity near the wall on the up-wash side
of the main vortex. As this accumulation grows throughout the cycle, the flow on the down-
wash side of the horseshoe vortex is entrained, moving it back towards the main vortex.

Another observation from the phase-averaged streamwise vorticity plots is the


spanwise convection of the main vortex during the actuation cycle. In order to aid the reader
in distinguishing this motion, a dashed-line has been included in Figs. 3.13b-g, tracking
the vortex cores through each phase. The main vortex convects the farthest in the spanwise
direction near the middle of the actuation cycle, corresponding to a peak vorticity in both
the positive vorticity of the main vortex and the negative vorticity near the wall. This
suggests that the effects of the interaction between the main vortex and the wall are not
limited to reducing the spanwise distance between the main and horseshoe vortices, but
also enhance the spanwise convection of the main vortex itself. This follows from the
previous discussion of representing the array of co-rotating vortices as four potential
vortices, explaining the spanwise convection of the pair by self-induction. With increased

33
vortex strength, there is an associated increase in the induced velocity experienced by the
vortex, which is clearly observed. Moreover, the interactions of opposite vorticity
contribute to the stretching of the main vortex, as was previously discussed when
comparing the time-averaged streamwise vorticity contour plots under static and dynamic
conditions. The stretching was more significant in the time-averaged static case than when
the VG was actuated. However, in the phase-averaged results, stretching of the main vortex
can also be observed in the static case. This is most prevalent in the beginning phases where
there is a region of negative vorticity present above the main vortex.

Comparing the time-averaged results for the static case and phase-averaged results
of the dynamic case, it can be observed that the vorticity concentration of both vortices is
greater during portions of the actuation cycle when the VG is dynamic. This suggests that
the strength, formation, and evolution of the vortices is highly susceptible to this dynamic
motion. It is hypothesized that the increased height of the VG during parts of the actuation
cycle results in stronger vortices. When the VG is taller in the free stream, it appears to the
flow as a larger obstruction, which can lead to enhancing the formation of the horseshoe
vortex. Expanding the parametric study of DVGs is required to determine the extent to
which mean height and amplitude effects the incoming boundary layer and subsequent flow
field.

Though vorticity fields are helpful in detecting velocity gradients and providing
explanation for certain flow behaviors, in order to better confirm the presence of vortical
structures, it is necessary to study Q-criterion plots. By definition, Q-criterion depict
regions that are dominated by rotation, i.e. a vortex. Figs 3.14a-e present a comparison
between the time- and phase-averaged iso-surfaces of Q-criterion in an effort to gain a
more comprehensive understanding of the evolution of these vortical structures over the
actuation cycle of the VG. Consistent to what was presented in the phase-averaged vorticity
contour plots, the trajectory of the vortices varies over the actuation cycle. During the
earlier phases (i.e.  = 0and 90), the main vortex is almost parallel with the free stream
velocity,  . The spanwise shift of the main vortex is most predominantly observed later
in the actuation cycle, at phases,  = 180and 270. Furthermore, the size of the main
vortex also varies as a function of the phase. The main vortex is smallest at the beginning
of the actuation cycle, as presented in Fig. 3.14a. It grows to its largest size during the

35
middle phases, which is shown in Figs. 3.14c-d. Finally, at the last phase angle presented,
the main vortex decreases in size. This suggests that the size of the main vortex is hightly
sensitive to the height of the VG, though there appears to be a delay in the response of the
flow field. At an input phase of  = 90, the VG is at its maximum height (i.e. h = 4.5 mm).
However, the largest effects on the flow field are experienced when  = 180. It is
hypothesized that the phase shift would vary as a function of free stream velocity, the length
of the VG, and its skew angle with respect to the free stream velocity.

Similar to the main vortex, the horseshoe vortex also varies as a function of phase angle.
During the first half of the actuation cycle, it is hardly present in the measurement domain.
It is unclear at  = 0 whether this is a result of the horseshoe vortex not forming, if it is
outside the measurement domain, or if the magnitude of rotation is below detection with
the current resoultion of the data. In the later half of the acutation cycle, the horseshoe
vortex can clearly be observed. Also consistent with the previously presented vorticity
contour plots are the differences in spanwise distance between the two vortices in the
middle of the actuation cycle (i.e.  = 180) and for later phases (i.e.  > 180). During the
latter portion of the cycle, the spanwise distance between the main and horseshoe vortices
is less than what is observed during the middle of the acutation cycle. The small pockets
of opposite rotation that exist between the two vortices, which are most prevalent at  =
180, is responsible for the spanwise shift of the horseshoe vortex towards the main vortex
as made evident during the examination of the vorticity contour plots.

36
3.2.3 Turbulent Kinetic Energy

The time-averaged results of the static and dynamic cases demonstrated that the
strength of the vortices present was greatest when the VG was not actuated. However, the
goal of the dynamic VG is to enhance flow mixing in order to best dissipate energy near
the wall surface. This would better ensure the delay of flow separation should dynamic
VGs be installed on the wings of aircraft. A method to measure enhanced mixing is first
through an analysis of the fluctuating velocity components, which consist of the random
turbulent fluctuations and the coherent fluctuations due to the actuation. The following
presented equations are given for the streamwise,  , velocity component only, however
the operations were also performed on the normal,  , and spanwise,  , velocity
components. The time-averaged velocity field can be divided into two components:

    (3)

where  is the acquired velocity,  is the mean velocity and  is the fluctuating
component or normal stresses, which consists of random turbulent fluctuations,  , and the
coherent fluctuations due to the actuation,  :

    (4)

The total normal stresses, acquired through equation 4, are used to calculate the
turbulent kinetic energy, TKE:


        ) (5)

  
where  ,   , and   are the total normal stresses.

The wall-normal distribution of TKE at the center of the main vortex for four
streamwise locations (i.e. x/h = 8, 12, 16, and 20) is shown in Figs. 3.15a-d. At all
streamwise locations along the height of the measurement domain, the TKE is consistently
higher for the dynamic case. Also present at these locations are two peaks, although they

38
4. Summary and Conclusions

Using oil flow visualization and SPIV, the interaction between a vortex generator and
a laminar boundary layer was investigated. The oil flow visualization experiments revealed
the presence of several flow structures, including the leading edge separation bubble, the
main vortex, the primary and secondary horseshoe vortices, and an area of induced
velocity. These results were helpful in determining the interrogation domain in which to
focus SPIV measurements and explaining the evolution of the vortical structures. Next, the
two- and three-dimensional time- and phase-averaged velocity fields were reconstructed
from multiple spanwise measurement planes. It was shown in the time-averaged results
that under both static and dynamic conditions, the main and primary horseshoe vortices are
present. Due to the location of the interrogation domain, it was not possible using the SPIV
measurement technique to confirm the presence of the leading edge separation bubble nor
the secondary horseshoe vortex detected in the oil flow visualization experiment. These
vortices act as a pair of co-rotating vortices, with their interaction being primarily
dependent upon the strength of the larger main vortex. It was observed that the strength of
the vortices was greatest immediately downstream of VG and decreased as a function of
downstream distance. However, as the main vortex developed farther downstream, there
was an accumulation of opposite vorticity on the up-wash side. As this grew, flow from
the down-wash side of the primary horseshoe vortex was entrained, causing it to move
closer to the main vortex, accelerating the vorticity dissipation of the horseshoe vortex. In
addition to the spanwise movement of the horseshoe vortex towards the main vortex, there
was an overall spanwise convection of the array of co-rotating vortices in the negative
spanwise direction. Using potential flow theory, it was shown that this convection was self-
induced with a velocity which varied as a function of the strength of the vortex and its
distance away from the wall. Furthermore, it was shown in the time-averaged results that
the interaction between the vortices and both and wall and outer flow ultimately resulted
in stretching the main vortex. This is due to the conservation of angular momentum.

Overall, comparing the time-averaged results for static and dynamic conditions, it was
shown that the strength of the vortices was greater when the VG was not actuated. However,
these results only depict the mean flow field and do not fully capture the evolution of the
vortices throughout the actuation cycle. Through the study of phase-averaged results,

40
where the flow field results have been phase-locked to the actuation of the VG, it is possible
to obtain a more in depth understanding of the flow physics associated with the dynamic
motion of the VG. First, it was shown that the size and location of the vortices not only
vary as a function of downstream distance, but also as a function of the phase. By the
middle of the actuation cycle, the strength of the vortices reaches a maximum. As a result,
the negative vorticity near the wall increases, drawing the horseshoe vortex closer to the
main vortex. However, on upstroke, the sudden increase in size and strength of the main
vortex initially forces the horseshoe vortex away. Due to the changing location and strength
of the vortices, the concentration of vorticity appears to be greater when the VG is static,
in the time-averaged flow field. Furthermore, as a result of varying size and location, the
vorticity associated with these vortices is dissipated over a greater area, leading to
enhanced flow mixing.

In order to confirm enhanced flow mixing, the turbulent kinetic energy, TKE, was
calculated in both the static and dynamic cases. It was shown from these results that there
was indeed an increase in TKE when the VG was actuated. Peaks in TKE were associated
with the vortices and other regions of vorticity concentrations. Moreover, the overall TKE
was higher for the dynamic as compared to the static case over the entire measurement
domain.

The goal of using DVGs is to enhance flow mixing. Though it has been shown that
stronger vortices are more efficient in transporting higher momentum fluid from the outer
flow into the boundary layer, it is equally important that the distribution of energy is well
dispersed in order maximize flow mixing. Comparing the static and dynamic cases, the
mean flow field revealed that the vortices were stronger when the VG was not actuated.
However, through the study of phase-averaged results and TKE, it was shown that the
dynamic motion of the VG was more effective in increasing flow mixing.

41
5. Future Work

The current work has focused on the interaction of a single vortex generator with a
laminar boundary layer of a flat plate. In this study, the measurement techniques of oil flow
visualization for a static VG and SPIV for both a static VG and DVG, were utilized. Future
work in this research includes drastically expanding this parametric study on a flat plate.
There are a number of parameters to be varied, including the shape of the VG. In this study,
the VG was rectangular in shape, with sharp corners. The most commonly used VGs that
are presently installed on aircraft are not designed in this manner, as they typically have a
leading edge curve. Another parameter that should be modified in the study of a single VG
in a cross-flow is the mean height of the VG. It has been shown for static vortex generators
that the effectiveness of a VG is dependent upon its height with respect to the local
boundary layer. Other parameters worth changing are the driving frequency and amplitude
of the DVG. It has been hypothesized that the dynamic motion of the DVG at certain
frequencies can excite frequencies already present in the flow, such as T-S waves.

Further expansion of this study includes studying an array of DVGs on an airfoil.


Static VGs are typically installed such that the resultant vortices from an array of VGs
interact with one another. This leads to an ever-growing parametric study. In this case, the
spanwise spacing and chordwise location of DVGs can be further explored. More extensive
use of the experimental techniques used in this study, along with additional techniques, i.e.
static and dynamic surface pressure measurements, should be used. While installed on an
airfoil model, the effect of the Reynolds number, angle of attack, and sweep angle are other
parameters worthy of future research.

42
6. References

Barth, Thomas, Scholz, Peter, and Wierach, Peter. "Flow Control by Dynamic Vane
Vortex Generators Based on Piezoceramic Actuators." AIAA Journal 49.5 (2011):
921-931. Print.

Casper, Jay, Lin, John, and Yao, Chung-Sheng. "Effect of Sub-Boundary Layer Vortex
Generators on Incident Turbulence." AIAA 33rd Fluid Dynamics Conference,
Orlando FL, 23-26 June 2003, AIAA Paper 2003-4162. Print.

Godard, G. and Stanislas, M. "Control of a Decelerating Boundary Layer. Part 1:


Optimization of Passive Vortex Generators." Aerospace Science and Technology
10.3 (2006): 181-191. Print.

Jenkins, Luther, Gorton, Susan, and Anders, Scott. "Flow Control Device Evaluation for
An Internal Flow with An Adverse Pressure Gradient." AIAA 40th Aerospace
Sciences Meeting & Exhibit, Reno NV, 14-17 January 2002, AIAA Paper 2002-
0266. Print.

Kerho, Michael and Kramer, Brian. "Enhanced Airfoil Design Incorporating Boundary
Layer Mixing Devices." 41st AIAA Aerospace Sciences Meeting & Exhibit, Reno
NV, 6-9 January 2003, AIAA Paper 2003-0211. Print.

Le Pape, Arnaud, Costes, Michael, Richez, Francois, Joubert, Gilles, David, F., and Deluc,
Jean-Michael. "Dynamic Stall Control Using Deployable Leading-Edge Vorex
Generators." AIAA Journal 50.10 (2012): 2135-2145. Print.

Lin, John. and Howard, Floyd. "Turbulent Flow Separation Control Through Passive
Techniques." 2nd Shear Flow Conference, Tempe AZ, 13-16 March 1989, AIAA
Paper 89-0976. Print.

43
Lin, John. "Control of Turbulent Boundary-Layer Separation Using Micro-Vortex
Generators." AIAA 30th Fluid Dynamics Conference, Norfolk VA, 28 June - 18
July 1999, AIAA Paper 99-3404. Print.

Lin, John, Robinson, Stephen, McGhee, Robert, and Valarezo, Walter. "Separation Control
on High LIft Airfoils via Micro-Vortex Generators." Journal of Aircraft 31.6
(1994): 1317-1323. Print.

Lin, John, Howard, Floyd, and Bushnell, Dennis. "Investigation of Several Passive and
Active Methods for Turbulent Flow Separation Control." AIAA 21st Fluid
Dynamics, Plasma Dynamic and Lasers Conference, Seattle WA, 18-20 June 1990,
AIAA Paper 90-1598. Print.

Lin, John, Howard, Floyd, and Selby, Gregory. "Small Submerged Vortex Generators for
Turbulent Flow Separation Control." Journal of Spacecraft and Rockets 27.5
(1990): 503-507. Print.

Lin, John, Selby, Gregory, and Howard, Floyd. "Exploratory Study of Vortex-Generating
Devices for Turbulent Flow Separation Control." 20th Aerospace Sciences
Meeting, 7-10 January 1991, Reno NV, AIAA Paper 91-0042. Print.

McCurdy, William. "Investigation of Boundary Layer Control of an NACA 16-325 Airfoil


by Means of Vortex Generators." United Aircraft Corp, Research Department,
Rept. M-15038-3 (3 December 1948). Print.

Osborn, Russell, Kota, Sridhar, Hetrick, Joel, Geister, Donald, Tilmann, Carl, and Joo,
Jinyong. "Active Flow Control Using High-Frequency Compliant Structures."
Journal of Aircraft 41.3 (2004): 603-609. Print.

Patel, Mehul, Carver, Reed, Lisy, Frederick, and Prince, Troy. "Detection and control of
flow separation using pressure sensors and micro-vortex generators." AIAA 40th
Aerospace Sciences Meeting & Exhibit, Reno NV, 14-17 January 2002, AIAA
Paper 2002-0268. Print.

44
Pearcey, H. H. Shock-induced separation and its prevention by design and boundary layer
control. Boundary Layer and Flow Control 2. Oxford: Pergamon, 1961. Print.

Perkins, Henry. "The formation of streamwise vorticity in tubulent flow." Journal of Fluid
Mechanics 44.4 (1970): 721-740. Print.

Rao, Dhanvada, and Kariya, Tsuyoshi. "Boundary Layer Submerged Vortex Generators
for Separation Control - an exploratory study." AIAA 1st National Fluid Dynamics
Conference, Cincinnati OH, 25 July 1988, AIAA Paper 88-3546. Print.

Schubauer, Galen and Spangenber, W.G. "Forced Mixing in Boundary Layers." Journal of
Fluid Mechanics 8.1 (1960): 10-32. Print.

Seshagiri, Amith, Cooper, Evan, and Traub, Lance "Effect of Vortex Generators on an
Airfoil at Low Reynolds Number." Journal of Aircraft 46.1 (2009): 116-122. Print.

Shabaka, I.M.M.A, Mehta, R.D., and Bradshaw, Peter. "Longitudinal Vortices Imbedded
in Turbulent Boundary Layers. Part 1. Single Vortex." Journal of Fluid Mechanics
155 (1985): 37-57. Print.

Sullerey, R.K., Mishra, S., and Pradeep, A.M. "Application of Boundary Layer Fences and
Vortex Generators in Improving Performance of S-Duct Diffusers." ASME 124
(2002): 136-142. Print.

Taylor, H. D. "The elimination of diffuser separation by vortex generators." United Aircraft


Corp, Research Department, Rept. R-4012-3 (10 June 1947). Print.

Vyas, Manan, Hirt, Stefanie, Chima, Rodrick, Davis, David, and Wayman, Thomas.
"Experimental Investigation of Vortex Generators on a Low-Boom Supersonic
Inlet." 29th AIAA Applied Aerodynamics Conference, Honolulu HI, 27-30 June
2011, AIAA Paper 2011-3798. Print.

45
Westerweel, Jerry. "Theoretical of the measurement precision in particel image
velocimetry." Experiments in Fluids 29 (2000): 3-12. Print.

46

Vous aimerez peut-être aussi