You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/312110373

Investigation of Crossflow Interaction of an Oscillating Jet

Conference Paper · January 2017


DOI: 10.2514/6.2017-1690

CITATIONS READS
21 886

7 authors, including:

Mohammad Arif Hossain Robin Prenter


The Ohio State University The Ohio State University
46 PUBLICATIONS 173 CITATIONS 21 PUBLICATIONS 152 CITATIONS

SEE PROFILE SEE PROFILE

Ryan Lundgreen Lucas Agricola


The Ohio State University The Ohio State University
13 PUBLICATIONS 92 CITATIONS 13 PUBLICATIONS 71 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fluidic Oscillator Heat Transfer View project

HPT Rotor Blade Tip Clearance Effect Mitigation View project

All content following this page was uploaded by Mohammad Arif Hossain on 18 December 2017.

The user has requested enhancement of the downloaded file.


10.2514/6.2017-1690
AIAA SciTech Forum
9 - 13 January 2017, Grapevine, Texas
55th AIAA Aerospace Sciences Meeting

Investigation of Crossflow Interaction of an Oscillating Jet

Mohammad A. Hossain1, Robin Prenter2, Ryan K. Lundgreen3, Lucas M. Agricola4,


Ali Ameri5, James W. Gregory6, Jeffrey P. Bons7
The Ohio State University, Columbus, OH, 43235

Numerical investigations were conducted to evaluate the interaction of an oscillating jet


with a crossflow. A conventional curved fluidic oscillator with an aspect ratio of unity is
used. The flow interaction is investigated at three different inclination angles to the crossflow
free stream direction ( = , , ) and at three blowing ratios ( = , , ). An
Unsteady Reynolds Averaged Navier-Stokes (URANS) model was used to evaluate the
flowfield. Models were validated by comparing key flow features and oscillating frequency
reported in the literature. Time averaged and time resolved flow fields are presented. Two
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

alternating streamwise vortices are observed at all blowing ratios. The sense of rotation of
these alternating vortices is opposite to the traditional counter rotating vortex pair (CRVP)
found for a steady jet in crossflow. A larger lateral jet spreading is observed for = at
higher blowing ratio. The trajectory envelope of a sweeping jet is shallower than the lowest
limit of the steady jet trajectory envelope for all cases due to rapid decay in velocity
magnitude. An investigation of the alternating vortices is presented at various phase angles
of the jet.

Nomenclature
AR = aspect ratio, (W/H)
BR = blowing ratio
CRVP = counter rotating vortex pair
D = hydraulic diameter, (10mm)
HSV = horseshoe vortex
HPV = hairpin vortex
RLVS = ring like vortical structure
SWV = streamwise vortices
Ut = throat velocity
U∞ = freestream velocity
= non-dimensional wall distance
= inclination angle
= streamwise vorticity, (1/s)
∅ = phase angle
Ω = vorticity magnitude
= strain rate tensor

I. Introduction

T he unsteady nature of an oscillating jet provides great potential for use as a flow control device, especially since
the frequency of oscillations is controllable by the designer. Since it involves no moving parts while generating
the oscillating jet, it is getting increased attention among the flow control community. Potential applications of such

1
Graduate Research Associate, Mechanical Engineering, Student Member AIAA.
2
Graduate Research Associate, Aerospace Engineering, Student Member AIAA.
3
Postdoctoral Researcher, Aerospace Research Center, Member AIAA.
4
Graduate Research Associate, Aerospace Engineering, Student Member AIAA.
5
Research Scientist, Mechanical and Aerospace Engineering.
6
Associate Professor, Aerospace Engineering, 201 W. 19th Ave, Associate Fellow AIAA.
7
Professor, Aerospace Engineering, 2300 West Case Road, Email: bons.2@osu.edu, Associate Fellow AIAA.
1
American Institute of Aeronautics and Astronautics

Copyright © 2017 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
a device include separation control1, drag reduction2, noise control3 and turbine film cooling4. Extensive research5-
6
has been performed on the development of active flow control technologies that force controlled reattachment of
separated flows over lifting surfaces, showing that aerodynamic performance can be improved by reattaching flow
with these devices. With pulsing, an effective separation control can be achieved while reducing the required
injected mass flow7-8 at the control surface. Two techniques that have been used successfully include synthetic jets,
zero-net-mass-flux devices that consist of enclosed cavities where alternating periods of blowing and suction are
created9 and high frequency plasma actuators. One of the main challenges in the design and development of active
flow control is a greater actuation power required to maintain effectiveness at full scale with higher free stream
velocities. Also, the majority of unsteady blowing actuators employ mechanical or piezoelectric devices which are
often large in size, and employ numerous moving parts with associated reliability issues and limited lifetime. For
this reason, in recent years there has been an increasing interest in fluidic actuator development10, which is a purely
passive system.
A significant amount of work has been done to understand the flowfield associated with jets in a crossflow.
Figure 1 shows the dominating flow structures of a transverse jet in a crossflow that includes shear layer vortices,
wake vortices, horseshoe vortices and a counter rotating vortex pair (CRVP). Shear layer vortices and wake vortices
are unsteady vortices that are strongest in the near field of the jet. Horseshoe vortices and the CRVP are more stable,
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

and convect downstream influencing the freestream11.

Figure 1. Different vortex system associated with the jet in a crossflow12.

The flow features of this fluidic oscillator have not been fully characterized especially the boundary layer
interaction and the effects of crossflow on the jet oscillation. Lacarelle et al13 showed improved mixing by four
fluidic oscillators in a generical jet in crossflow configuration compared to standard non oscillating jets. Ostermann
et al14 have studied the internal and external flow features of a fluidic oscillator in a quiescent environment using
phase averaged Particle Image Velocimetry (PIV). Later Osterman et al15 reported time-resolved PIV data and three
dimensional flow field for the same device in a crossflow environment where they studied a single blowing ratio
(BR = 3) at a single inclination angle ( = 90 ). According to their findings, oscillating jet is bent by the crossflow
and the trajectory of the sweeping jet is shallower than that of a steady jet in crossflow. In addition, spatial
oscillation creates a pair of counter-rotating vortex system and the sense of rotation is opposite to that of the counter-
rotating vortex pair of a steady jet in crossflow.
The primary goal of this study is to investigate the interaction between an oscillating jet in a cross flow at
different inclination angles and various blowing ratios where blowing ratio is defined by –

( )=

The fluidic oscillator used in this study was invented by Bowles Fluidic Corporation16. It is a conventional
feedback type fluidic oscillator which includes a fluidic circuit. The fluidic circuit consists of a power nozzle, two
feedback loops and an exit aperture, or throat. Figure 2 shows the schematic of a typical feedback type fluidic
oscillator. The general working principle of such a device is as follows. A jet enters into the cavity from a
pressurized plenum via the power nozzle. Then the power jet expands to fill the throat and the feedback channels.
Two opposite vortices begin to form on both sides of the power jet. As the intensity of the vortices increase, one

2
American Institute of Aeronautics and Astronautics
vortex becomes dominant. This causes the power stream to deflect against the opposite wall. When the power stream
deflects to the side wall, it attaches to the wall according to the Coanda effect. This allows a portion of the fluid to
enter into the feedback loop. This fluid carries pressure waves back to the control port, which helps the power
stream detach from the side wall. The power stream then switches to the opposite wall and the same process repeats,
which results in an oscillatory fluid motion at the throat.
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 2.Schematic of a feedback type fluidic oscillator.

II. Numerical Investigation

Computational Domain and Grid generation


Three different geometric configurations were developed based on the inclination angle ( ) of the oscillator with
the main stream. Figure 3 shows each geometry, with the fluidic oscillator and the direction with the freestream. The
exit of the oscillator was modified based on the inclination angle ( ). The exit opening area of the oscillator
increases as α becomes shallower with the streamwise flow direction. At = 90 , the exit section has a rectangular
opening which is extended 3.5D in the spanwise (z-axis) direction. For = 60 , the exit opening becomes a
trapezoidal shape. The exit area is approximately 90% of the actual rectangular opening used for = 90 . However,
the distance between the windward and the leeward edge remains constant (1D). For = 30 , the exit opening is
approximately 220% larger than the actual rectangular opening ( = 90 case). In addition, the distance between
the windward and the leeward edge is approximately 2D for this configuration. Figure 4 shows the computational
domain which is extended 40D in the streamwise direction and 20D (±10 from the centerline) in the spanwise
direction. The domain is extended 15D in the vertical direction (y-axis).The main flow inlet is located 10D upstream
of the oscillator windward edge. Hybrid grids including prism layers at the wall and polyhedral cells in the core of
the domain were used. Three grids were examined for this study. Table 1 shows the properties of each grid in terms
of a non-dimensional wall distance, .

Figure 3.Geometric configurations and oscillator exit geometry.


3
American Institute of Aeronautics and Astronautics
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 4.Computational domain and grid generation.

Table 1.Details of the computational grid and cases

CFD BR Cells y+ Time step


Case (million) Δt,(Sec)
1 2.42 4.75 2.0E-04
α = 90o 2 2.42 4.75 1.0E-04
3 2.42 4.75 5.0E-05
1 2.69 3.25 2.0E-04
α = 60o 2 2.69 3.25 1.0E-04
3 2.69 3.25 5.0E-05
1 2.74 3.50 2.0E-04
α = 30o 2 2.74 3.50 1.0E-04
3 2.74 3.50 5.0E-05

Model Validation
An unsteady RANS (URANS) simulation was performed in this study using the commercial code FLUENT to
predict the time averaged and time resolved flow fields. The flow equations were discretized by a second order
upwind scheme and spatial gradients were reconstructed by a least square cell based method. The k-ω SST model by
Menter et al.18 was employed to model the effects of turbulence. A second order implicit discretization in time has
been adopted for the unsteady calculation. A velocity inlet ( = 15 / ) boundary condition was assumed at the
main inlet and a massflow inlet condition was assumed at the inlet of the oscillator. The massflow was calculated
based on the theoretical throat velocity of the oscillator. The main flow inlet turbulence intensity was 0.4%, and the
oscillator inlet turbulent length scale was 0.01m. The outlet temperature and pressure were assumed ambient.
The model is validated with experimental results reported by Ostermann et al.15. Since the flowfield of the fluidic
oscillator is inherently unstable, predicting the frequency of the oscillation is important. The oscillation frequency of
the fluidic device reported by Ostermann et al.15 was 67 Hz at BR = 3, while the model predicted an oscillation
frequency of 71Hz at BR = 3, within 5% of the actual frequency. Figure 5 shows experimental and computational
instantaneous velocity vectors for BR = 3 in a crossplane at x/D = 14. The contours are colored by streamwise
velocity normalized by freestream velocity. The left column shows the experimental results reported by Ostermann
et al.15 at three phase angles of the jet, and the right column shows the data from this model. Many similarities were
observed in the velocity fields. The CFD prediction of jet spreading in the spanwise direction and jet penetration
height is qualitatively similar to the experimental results. In addition, the location and the size of the vorticies are
similar compared to experimental results. The sign of the vorticity is also matched in both cases. It is important to
note that the boundary layer profile reported by Ostermann et al.15 was fully turbulent at the exit of the oscillator.
However, a uniform velocity inlet condition was used for the CFD in this study. The boundary layer thickness ( )

4
American Institute of Aeronautics and Astronautics
was approximately 15mm (3D) in the experimental case while in CFD, the boundary layer thickness was
approximately 5mm (0.5D) at the leeward edge of the oscillator exit.
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 5.Instantaneous crossplane velocity vectors overlaid on contours of normalized streamwise velocity at
x/D = 14, BR = 3 and =

Another qualitative validation was made by comparing the 3D flowfield at BR = 3. Figure 6 shows the time
resolved isosurface of Q = 130,000 s-2 for four different phase (∅) angles with data from Ostermann et al.15 and the
current model. Here Q is defined as a function of vorticity magnitude (Ω ) and strain rate sensor ( ). (Eq. 1)

1
= Ω − (1)
2

The presented phase angles of ∅ = 0 , 45 , 90 and135 make up one “sweep” of the jet, starting attached to one
side of the oscillator and switching to the other. Due to symmetry, the data for ∅ = 180 , 225 , 270 and 315 is
theoretically identical, in the reversed direction. It is evident that CFD was able to predict the dominant flow
structures such as the alternating streamwise vortices and their directions. However, some small structures are also
predicted by the current CFD model that are not seen in the experimental data. The primary reason for this is that the
PIV data presented by Ostermann et al.15 are phase-averaged which involves spatial smoothing. The phase-
averaging has 2 dominant effects: 1) if the vortex appears in a slightly different place each time, the averaging will
make it look more diffuse than it really is in a single snap-shot and 2) random smaller vortices that spin off from the
larger structure probably appear at different places each time and will be lost in the phase averaging. There’s also
spatial smoothing in any PIV system due to the minimum box size in the processing and any smoothing that was
used to get rid of spurious vectors. In contrast, the CFD data presented here is essentially (is it, or not) time accurate.
In addition, the current RANS turbulence model does not have enough dissipation to transfer energy from large
structures to relatively small structures. The term “coherent structure” is used to describe the dominant streamwise
vortex (SWV) structures in the flow field downstream of the fluidic oscillator, before they mix out into the main

5
American Institute of Aeronautics and Astronautics
flow. These large coherent structures are observed in the near wall region for all cases. In addition, some other
important flow structures such as ring-like vortices (RLVS) and horseshoe vortices (HSV) are also visible in the
CFD results and identified in Fig. 6.
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 6. Instantaneous flow field (top view) at various phase angle(∅). Isosurface of Q = 130,000s-2 colored
by streamwise vorticity at BR = 3 and  = 90°.

III. Results

This section discusses the time averaged velocity field of the fluidic device for three different inclination angles
( = 90 , 60 and 30 ) and three blowing ratios (BR = 1,2,3). In addition, the time resolved vortical structures are
presented at four different phase angles (∅ = 0 , 45 , 90 and 135 ). Figure 7 shows the oscillation frequency of
the jet at the exit of the device. As expected, the oscillation frequency increases with the increase of blowing ratio.
Since the sweeping mechanism is dominated by the internal geometry of the oscillator, the frequency of the jet does
not change when it interacts with the cross stream flow. The frequency estimated by CFD is also compared with
experimental frequency reported by Ostermann et al.15 for a similar curved feedback type fluidic oscillator. A
slightly higher frequency (71 Hz at BR = 3) was predicted by CFD compared to the experimental frequency (67 Hz).

100

80
Frequency(Hz)

60

40
0
CFD (  = 30 )
0
CFD (  = 60 )
20 0
CFD (  = 90 )
0
Exp (  = 90 ) (Ostermann et al. 2016)

0
0 2 1 3 4
Blowing ratio (BR)
Figure 7. Oscillation frequency at various blowing ratios.

6
American Institute of Aeronautics and Astronautics
Figure 8 shows the time-resolved flow field (top view) for = 90 case at three blowing ratios for half an
oscillation period. The flow field is presented by isosurfaces of = 220,000 colored by streamwise vorticity.
Each column of Figure 7 represents a single oscillation phase (∅) at various blowing ratios and each row shows a
sequence of phases for half an oscillation. As one might expect, two alternating streamwise vortices were found as
the most dominant flow structure. The alternating behavior of the vortices is generated by the sweeping action of the
jet interacting with the main flow. The sense of rotation of these two streamwise vortices are opposite from each
other, which was also reported by Ostermann et al.15. It is important to note that the sense of rotation of these
streamwise vortices is opposite to the counter rotating vortex pair (CRVP) found in a steady jet in crossflow. The
strength of these streamwise vortices decreases due to the sweeping action of the jet which enhances mixing as they
convect downstream. It was also observed that the flow field is most dynamic near the leeward side of the hole. As
soon as the jet begins interacting with the mainflow, it starts to break down into smaller structures that augment
mixing. With the sweeping action of the jet, this mixing can cover a larger surface area making it attractive for
applications such as film cooling. A noticeable ring-like vortical structure was observed at the highest blowing ratio
(BR = 3) where the jet penetrates the furthest from the wall. These ring-like vortices are also common in transverse
jets in crossflow17. However, it is not well understood how they interact with the two alternating streamwise
vortices.
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 8. Instantaneous flow field (top view) for = configuration at various phase angles(∅) and
blowing ratios (BR = 1-3). Isosurface of Q = 220,000s-2 colored by streamwise vorticity.

The effects of blowing ratio on the flow field are also shown in Figure 8. As blowing ratio decreases (BR = 2 and
1), the lateral spreading of the jet also decreases. This happens due to lack of momentum required to penetrate into
the main flow. However, the alternating vortices still dominate the flow field as they convect downstream. The ring-
like vortices (RLVS) appeared slightly ( = 2 , ∅ = 0 and 90 ), but did not continue with the streamwise
vortices. Strong mixing was not observed in these cases compared to BR = 3 case. At the lowest blowing ratio (BR
= 1) the jet spreading was the least and horseshoe vortices and hairpin vortices were visible at this blowing ratio.
Sau and Mahesh17 reported the presence of these hairpin vortices at low blowing ratios in a steady jet case. More
details of the jet spreading mechanism are explained with streamlines later in this section.

7
American Institute of Aeronautics and Astronautics
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 9. Instantaneous flow field (top view) for = configuration at various phase angles(∅) and
blowing ratios (BR = 1-3). Isosurface of Q = 220,000s-2 colored by streamwise vorticity

Figure 10. Instantaneous flow field (top view) for = configuration at various phase angles(∅) and
blowing ratios (BR = 1-3). Isosurface of Q = 220,000s-2 colored by streamwise vorticity

Figure 9 and 10 show a similar time-resolved flow field (top view) for = 60 and 30 cases at three blowing
ratios for half an oscillation period. The flow fields are also presented by isosurface of = 220,000 colored by
streamwise vorticity. The exit opening of the device changes, as the inclination angle becomes shallower ( =
8
American Institute of Aeronautics and Astronautics
60 and 30 ). The exit opening takes a trapezoidal shape from a rectangular shape. Two alternating streamwise
vortices are visible. No significant variation was observed in jet spreading since the oscillation frequency and
blowing ratio remain the same in these cases and all other parameters such as boundary layer thickness and
freestream velocity remain constant. With similar blowing ratios, the momentum carried out by the jet is similar for
both cases that are responsible for the jet spreading. However, there are some noticeable differences in coherent
structures. For = 60 case, the coherent structures survive a larger streamwise distance compared to = 90 . As
seen in the previous case ( = 90 , = 3 , ∅ = 0 ), the coherent structures cover approximately 8D in the
streamwise direction. In contrast, these structures remain coherent approximately 11D in the streamwise direction
for this ( = 60 , = 3 , ∅ = 0 ) case. A similar trend was also observed for = 30 case where the coherent
structures cover 14D along the streamwise direction (Figure 10). A lower convective velocity near the wall due to
boundary layer and higher streamwise velocity due to hole inclination is responsible for these structures to remain
coherent for a longer period of time. The ring-like vortices (RLVS) were observed at BR = 3 for = 60 . However,
no RLVS were observed for = 30 case as the inclination angle is such that the jet is not penetrating enough to
generate those RLVS.
Figure 11-12 show the instantaneous streamlines of all three cases at three blowing ratios released from the point
of maximum velocity at the throat. The streamlines presented here are colored by instantaneous phase angle of the
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

jet. Since the jet is oscillating spatially, the streamlines will change with time. Figure 11 shows the top view of the
instantaneous streamlines where each column of the figure show streamlines at different blowing ratios (BR) and
each row show the streamlines at different inclination angles ( ). It is evident that the streamlines are bent towards
the streamwise direction at the far field (x/D > 5) due to the presence of a wake behind the oscillating jet and
oppositely signed streamwise vortices. However, the sweeping of the jet causes the near-hole streamlines to deflect
and spread laterally. The maximum spreading is restricted by the freestream flow compared to the jet spreading in a
quiescent environment. Therefore, high blowing ratio causes higher lateral spreading (Fig.11 first column) of the jet
due to its high momentum compared to the local freestream momentum. During a single oscillation cycle the
streamline exhibits significant bending in the spanwise direction when the jet has minimum defection (∅ ≈ 90 ).
Ostermann et al.15 suggested that this bending happens due the presence of a local vortex trailing the jet during the
sweeping motion. The bending pattern of the streamlines changes with different inclination angle ( ). For = 90 ,
the maximum bending of the streamline occurs approximately at x/D = 2 while this bending location moves further
downstream from the hole as decreases ( = 60 , 30 ). As previously discussed, the vortical structures remain
coherent at = 60 , 30 for a longer period of time and the bending of the streamlines is delayed.
Figure 12 shows the jet trajectories by instantaneous streamlines of all three cases at three blowing ratios
released from the point of maximum velocity at the throat. As discussed earlier, the streamlines presented here are
colored by instantaneous phase angle of the jet. Each column of the figure show streamlines at different blowing
ratios (BR) and each row show the streamlines at different inclination angle ( ). These jet trajectories are compared
with the envelope of steady jet trajectory available in the literature12. It is evident that the streamlines are shallower
than the lowest limit of the steady jet trajectory envelope. This happens due to the sweeping action of the jet that
causes a rapid decay in velocity magnitude14. Moreover the streamline bends toward the wall due to the streamwise
vortices at the downstream edge of the hole. As the blowing ratio (BR) decreases, the streamlines become shallower
and remains closer to the wall. Similar behavior was observed for the = 60 , 30 case. Figure 12 also shows the
jet penetration height for each case. It is evident that the maximum jet penetration occurs at the highest blowing ratio
for a corresponding . As the blowing ratio decreases, the jet penetration height is also reduced due to the lack of
momentum. For = 90 , the estimated maximum jet penetration was 4D from the wall. The jet penetration height
reduces 100% for this case when the blowing ratio reduces from 3 to 1.
Figure 13 shows the time averaged velocity vectors in a crossplane at x/D = 14 colored by streamwise vorticity.
Some key features were observed from these time averaged results. The dominant alternating vortices are visible for
all cases and the sense of rotation is opposite to the CRVP found in a steady jet in crossflow. The lateral jet
spreading is also visible in Fig. 13. The highest jet spreading was found at = 90 , = 3 which is approximately
8D (±4 from the centerline).This lateral jet spreading is reduced significantly as the blowing ratio
decreases ( = 1). Moreover, the strength of vortices decreases significantly compared to the steady jet due to the
sweeping action of the jet that causes a rapid decay in velocity magnitude. This might be advantageous for
applications such as film cooling where relatively weak streamwise vortices are required to prevent the hot flow
from penetrating close to the cold coolant flow.

9
American Institute of Aeronautics and Astronautics
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 11.Instantaneous streamlines (xz-plane) colored by phase angle releases from the throat for =
, , configuration at various blowing (BR) ratios.

Figure 12.Instantaneous streamlines (xy-plane) colored by phase angle releases from the throat for =
, , configuration at various blowing (BR) ratios.

10
American Institute of Aeronautics and Astronautics
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

Figure 13.Time averaged velocity vectors in a crossplane at x/D = 14 colored by streamwise vorticity.

IV. Conclusion

Unsteady RANS calculations were conducted to evaluate the crossflow interaction of a sweeping jet emitting at
various inclination angles. A conventional curved fluidic oscillator with an aspect ratio of one was studied at
several blowing ratios. The time averaged and time resolved velocity fields were examined to understand the
development of the dominant flow structures. Oscillation frequency, jet spreading and jet penetration were
estimated at several blowing ratios. Some key findings are listed below-

1. The oscillation frequency of the jet does not change with interaction of the crossflow. The jet lateral
spreading and jet penetration height depends on the blowing ratio.
2. High blowing ratio improves mixing and a faster decay of the jet. Two alternating streamwise
vortices dominate the flow. In addition, high blowing ratio flows exhibit important flow structures
such as ring-like vortices and horseshow vortices.
3. The trajectory envelope of a sweeping jet is shallower than the lowest limit of the steady jet
trajectory envelope for all cases due to rapid decay in velocity magnitude.
4. The streamline bends toward the wall due to the streamwise vortices at the downstream edge of the
hole. As the blowing ratio (BR) decreases, the streamlines become shallower and remain close to the
wall.

Future studies will investigate the effect of the incoming boundary layer, streamwise pressure gradient and rate of
entrainment.

Acknowledgement

This work has been funded by the US Department of Energy (DOE - NETL) under award no. DE-FE0025320 with
Robin Ames as program manager. Computational resources were provided by the Ohio Supercomputer Center
(OSC). The views expressed in the article are those of the authors and do not reflect the official policy or position of
the Department of Energy or U.S. Government.

11
American Institute of Aeronautics and Astronautics
References
1
Cerretelli, C. and Kirtley, K., “Boundary Layer Separation Control with Fluidic Oscillators,” Journal of Turbomachinery, Vol.
131, No. 4, 2009. doi:10.1115/1.3066242.
2
Woszidlo, R., Stumper, T., Nayeri, C., and Paschereit, C. O., “Experimental Study on Bluff Body Drag Reduction with Fluidic
Oscillators,” 52nd AIAA Aerospace Sciences Meeting, Jan 2014.doi:10.2514/6.2014- 0403.
3
Raman, G. and Raghu, S., “Miniature fluidic oscillators for flow and noise control - Transitioning from macro to micro fluidics,”
AIAA Fluids 2000 Conference and Exhibit, 2000. doi:10.2514/6.2000-2554.
4
Thurman D, Poinsatte P, Ameri A, Culley D, Raghu S, Shyam V. Investigation of Spiral and Sweeping Holes. ASME. J.
Turbomach. 2016;138(9):091007-091007-11. doi: 10.1115/1.4032839
5
Tilmann, C.P., Kimmel, R.L., Addington, G.A., and Myatt, J.H., “Flow Control Research and Applications at the AFRL’s Air
Vehicles Directorate,” AIAA Paper 2004-2622, 2004
Downloaded by Tonya Maniaci on December 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-1690

6
Lin, J.C., Howard, F.G., Bushnell, D.M., and Selby, G.V., “Investigation of Several Passive and Active Methods for Turbulent
Flow Separation Control,” 21st Fluid Dynamics, Plasma Dynamics and Lasers Conference, AIAA, Washington, D.C.
7
Seifert, A. and Pack, L.G., “Active Control of Separated Flows on Generic Configurations at High Reynolds Numbers,” AIAA
99-3403, June – July 1999
8
Wygnanski, I. “Boundary Layer and Flow Control by Periodic Addition of Momentum,” AIAA Paper 97-2117.
9
Greenblatt, D. & Wygnanski, I., “The control of Flow Separation by periodic addition of momentum,” Progress in Aerospace
Sciences, Vol. 36, pp.487-545, 2000
10
Amitay, M., Smith, B.L., and Glezer, A., “Aerodynamic flow control using synthetic jet technology,” AIAA Paper 98-0208
11
Mahesh, K., “The Interaction of Jets with Crossflow,” Annual Review of Fluid Mechanics, Vol. 45, No. 1, 2013, pp. 379–407
doi:10.1146/annurev-fluid-120710-101115
12
Margason RJ. 1993. Fifty years of jet in cross flow research. Comput. Exp. Assess. Jets Cross Flow, AGARD-CP
534, Advis.Group Aeronaut. Res. Dev., Washington, DC
13
Lacarelle, A. and Paschereit, C. O., “Increasing the Passive Scalar Mixing Quality of Jets in Crossflow With Fluidics
Actuators,”Journal of Engineering for Gas Turbines and Power, Vol. 134, No. 2, 2012, pp. 21503. doi:10.1115/1.4004373.
14
Ostermann, F., Woszidlo, R., Nayeri, C.N. and Paschereit, C.O., “Experimental Comparison between the Flow Field of Two
Common Fluidic Oscillator Designs,” 53th AIAA Aerospace Sciences Meeting and Exhibit, Jan 2015. AIAA 2015-0781
15
Ostermann, F., Woszidlo, R., Nayeri, C.N. and Paschereit, C.O., “The Time-Resolved Flow Field of a Jet Emitted by a Fluidic
Oscillator into a Crossflow,” 54th AIAA Aerospace Sciences Meeting and Exhibit, Jan 2016. AIAA 2015-0345
16
Stouffer, R., “Oscillating spray device,” Patent US 4151955, 1979.
17
Sau R, Mahesh K. 2008, “Dynamics and mixing of vortex rings in crossflow,” J. Fluid Mech. 604:389–409
18
Menter, F. R., 1994, “Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications,” AIAA Journal, vol. 32,
No. 8, 1994, pp. 1598-1605

12
American Institute of Aeronautics and Astronautics

View publication stats

You might also like