You are on page 1of 16

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

net/publication/266387500

The Time-Resolved Internal and External Flow Field Properties of a Fluidic


Oscillator

Conference Paper · January 2014


DOI: 10.2514/6.2014-1143

CITATIONS READS
27 404

5 authors, including:

Rene Woszidlo Florian Ostermann


The Boeing Company Technische Universität Berlin
51 PUBLICATIONS 772 CITATIONS 23 PUBLICATIONS 294 CITATIONS

SEE PROFILE SEE PROFILE

Christian Navid Nayeri Christian Oliver Paschereit


Technische Universität Berlin Technische Universität Berlin
169 PUBLICATIONS 1,647 CITATIONS 622 PUBLICATIONS 6,637 CITATIONS

SEE PROFILE SEE PROFILE

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

Shockless Explosion Combustion View project

Collaboration with TU Berlin on quiet WT blades View project

All content following this page was uploaded by Florian Ostermann on 16 March 2016.

The user has requested enhancement of the downloaded file.


The Time-Resolved Internal and External Flow Field
Properties of a Fluidic Oscillator

S. Gärtlein∗, R. Woszidlo†, F. Ostermann∗,


C.N. Nayeri‡, and C.O. Paschereit§
—Hermann-Föttinger-Institut—
Technische Universität Berlin, 10623 Berlin, Germany

Although the effectiveness of fluidic oscillators as flow control actuators is well known, a
lack of knowledge regarding their internal mechanisms and external dynamics persists. Due
to their commonly small size and high oscillation frequency, time-resolved measurements
are challenging. In this study an enlarged fluidic oscillator made out of acrylic glass and
supplied with air under pressure is investigated experimentally. Combined with a high-
speed PIV system and time-resolved pressure measurements inside the fluidic oscillator,
high temporal resolution is accomplished. The results reveal the underlying mechanism
causing the jet to oscillate. The knowledge of the internal dynamics allows an educated
improvement and optimization of the oscillator’s design. Several properties (e.g., deflection
angle, jet width, jet velocity, and entrainment) of the oscillating jet, which is emitted into
a quiescent environment, are examined. Results show that the jet’s characteristics are
only independent of the supply rate within a limited range. Although well within the
incompressible regime, the jet’s oscillation pattern changes at higher supply rates. The
time-resolved flow field infers that the oscillating jet resides longer in its deflected state
than it takes to switch over to the other side. Throughout one oscillation cycle the jet’s
properties are also found to oscillate substantially which may be of significance for various
applications. A comparison with a free axisymmetric jet reveals that the oscillating jet
entrains more fluid. This behavior is accompanied by a faster decrease in jet velocity and
increase in jet width.

I. Introduction
luidic oscillators generate a self-induced and self-sustaining spatially oscillating jet when supplied with
F fluid under pressure. Some of the major advantages they have over other active flow control actuators
include their simplicity and practicability, as the oscillation only depends on the oscillator’s geometry, size,
and flowrate without the need for any moving parts. Fluidic oscillators, which were already developed
more than half a century ago, have mainly been used with water as the working fluid.1 In recent years the
interest in fluidic oscillators as actuators for flow control has increased noticeably. As a result, numerous
studies have been published investigating the effectiveness of fluidic oscillators. Some remarkable results were
produced, especially in separation control.2, 3, 4, 5 However, fluidic oscillators were also used for bluff body
drag reduction,6, 7 noise reduction,8, 9 and combustion control.10
While the effectiveness of fluidic oscillators is evident, the reasons for their efficiency and the underlying
mechanism remain unclear. There are few studies investigating the oscillator’s switching mechanism and
flow field experimentally11, 12, 13, 14 and numerically.15 Although, these studies provide an insight into the
dynamics of fluidic oscillators, all experimental investigations employed some simplifications. Either water
was chosen as the working fluid12, 13 or an external trigger was used.14 Due to the compressible nature of air,
different effects are expected, particularly at high Mach numbers. External triggering suppresses the jet’s
∗ Graduate Student, Hermann-Föttinger-Institut, Technische Universität Berlin.
† Postdoctoral Scholar, Hermann-Föttinger-Institut, Technische Universität Berlin, AIAA Member.
‡ Research Assistant, Hermann-Föttinger-Institut, Technische Universität Berlin, AIAA Senior Member.
§ Professor, Chair of Fluid Dynamics, Hermann-Föttinger-Institut, Technische Universität Berlin, AIAA Senior Member.

1 of 15

American Institute of Aeronautics and Astronautics


natural oscillation possibly resulting in different dynamics. As a result, numerical investigations do not have
any reliable data for validation. Hence, an uncertainty of previous numerical studies remains. Ostermann et
al.16 describe the difficulties of experimentally acquiring time-resolved data from small-scale devices and
high-frequency naturally oscillating flow fields. Experimental and analytical methods were established to
obtain time-resolved data for fluidic oscillators even at high frequencies and without an external trigger.
In this study the phase-averaging method suggested by Ostermann et al.16 is applied on an enlarged fluidic
oscillator supplied with pressurized air and without an external trigger. The internal flow field reveals the
switching mechanism and internal dynamics of fluidic oscillators, yielding several suggestions for possible
design improvements. Further, the external flow field is examined, which infers substantially differing
properties when compared to a steady axisymmetric jet. The study also provides a data set for validation of
numerical studies.

II. Setup and Instrumentation


The fluidic oscillator investigated in this study is shown in Fig. 1. For the initial experiments the oscillator
is enlarged appropriately to reduce the oscillation frequency for air as a working fluid. Therefore, the
phase-averaging methods are validated by high-speed PIV measurements. The frequency is consequently
reduced from the order of several kHz to the order of 10 Hz. The enlarged oscillator is machined out of
acrylic glass with the cavities being 25 mm deep. Another acrylic glass plate covers the oscillator and provides
optical access. An airtight seal is ensured throughout the experiments. The fluidic oscillator is supplied with
compressed air, which is controlled by a digital massflow meter. The plenum is separated from the oscillator
inlet by a section of honeycomb guaranteeing homogeneous inflow conditions. Furthermore, symmetrical
boundary conditions and unobstructed outflow conditions are ensured by the given setup (Fig. 2).

outlet

outer
region

inlet mixing chamber


plenum

feedback channel
(FBC)

FBC outlet FBC inlet

Figure 1. Examined fluidic oscillator with denotations of main regions.

For Particle Image Velocimetry (PIV), a high-speed camera (Photron Fastcam SA1.1) and a Nd:YLF Laser
(Quantronix Darwin Duo 100) with a mirror arm and light sheet optics are used. The mirror arm is positioned
adequately from the oscillator’s exit and the camera is positioned perpendicular to the laser sheet as depicted
in Fig. 2. An aerosol generator is added to the air supply system for seeding. For measurements of the
external flow field, a second aerosol generator emits seeding into the surrounding air prior to the experiments.
The sampling rate of 1500 Hz is two orders of magnitude larger than the oscillation frequency, ensuring a
high temporal resolution. As a naturally oscillating flow with fluctuating frequencies is investigated, the PIV
sampling rate is set to a constant value. 10,920 PIV double pictures are recorded with a resolution of one
megapixel. The distance of the camera is chosen to yield a spatial resolution of approximately four pixels per

2 of 15

American Institute of Aeronautics and Astronautics


millimeter. The pulse distance of the camera pictures are chosen according to the maximum velocity. The
final interrogation window is set to 12 × 12 pixels in the inner flow field and 16 × 16 pixels for the external
flow with an overlap of 50%. This corresponds to a spatial resolution of approximately 1.5 mm for the inner
flow field and 2 mm for the outer flow field. Since air is used as the working fluid no refractive index matching
is possible. Therefore, parts of the geometry act as prisms. Consequently, the illumination is not homogenous
inside the actuator. Therefore, three different laser positions are applied to fully resolve the internal flow
field. During postprocessing, the phase-averaging method based on the simultaneously recorded pressure
signal inside the oscillator allows the reconstruction of the entire internal flow field.16 For the external flow
field, four camera positions are chosen in order to extend the measured area while maintaining the spatial
resolution.
fluidic oscillator
with pressure taps

honeycomb

air supply camera

laser

Figure 2. Schematic of the general setup.

Inside the oscillator 55 pressure taps are distributed (Fig. 2) according to the study by Bobusch et al.12
The pressure range of the transducers (HDO Series by Sensortechnics) is chosen appropriately for the local
flow conditions. Their sampling rate is three orders of magnitude larger than the oscillation frequency. The
size of the pressure taps is optimized regarding minimal disturbance of the flow and phase delay as well as
resonance effects in the region of interest employing the calculation for the dynamic response of tubes.17
All transducers are recorded simultaneously with a multichannel DAQ system from National Instruments.
Additional information on the setup and instrumentation is provided by Ostermann et al.16

III. Data Analysis


The described setup ensures time-resolved measurements as the sampling rates of all instrumentations are
considerably larger than the oscillation frequency. Ostermann et al.16 established a phase-averaging method
for naturally oscillating flow fields and conventional PIV sampling rates. The presented data in this paper is
phase-averaged over a window size of 3◦ using this method. The general phase-averaging steps are:
1. A reference signal has to be found that clearly exhibits the oscillation. In this study, the difference of
two symmetrically positioned pressure sensors in the feedback channel inlet demonstrates remarkably
good qualities.

3 of 15

American Institute of Aeronautics and Astronautics


2. A numerical low pass filter is applied forward and backward to reduce noise while ensuring the
conservation of phase and amplitude.
3. An autocorrelation is applied on the reference signal with a segment length of one half to one oscillation
period length. The start of one oscillation period is defined as the sign change from negative to positive
values of the time-resolved correlation coefficient.

4. The identified periods are used to assign a phase angle to each PIV snapshot.
5. The snapshots are averaged within a specific phase angle window size which is based on minimized
RMS values.
As explained in Chapter II, several laser and camera positions are applied. Hence, these PIV fields have to
be merged and phase-aligned. Therefore, the phase lag of the phase-averaged reference signals is identified
and the fields shifted accordingly. A smooth transition is achieved by weighted averages between the original
fields. Furthermore, a global definition of the starting point of an oscillation cycle is required for a proper
comparison between different test cases. The literature does not offer any quantitatively based definition.
In fact, it appears that phase angle assignments found in other studies are neither coherent nor bare any
quantitative criterion. Most discussions are based on a qualitative criterion when the jet is fully deflected
internally or externally. However, the criterion is challenging to quantify and strongly dependent on the
internal or external location. Therefore, the global phase angle start in this study is defined as the zero
pressure difference between the upper and lower feedback channel (FBC) inlet with a sign change from positive
to negative. Qualitatively, this marks the moment when the recirculation bubble at the upper wall has its
smallest size and the jet exits the oscillator at almost zero deflection. However, these qualitative observations
are strongly dependent on the specific oscillator design and size, which affect the internal convection velocities
and distances.
The employed method yields one continuous phase-averaged internal and external flow field for various supply
rates. The time-resolved information allows a detailed examination of the internal dynamics and jet properties
(e.g., deflection angle, jet width, and entrainment). For structural clarity, the internal and external flow fields
are discussed separately in the following chapter.

IV. Results
The investigation of the fluidic oscillator is conducted with numerous supply rates ranging up to 100 kg/h.
The corresponding Reynolds numbers are based on the exit’s hydraulic diameter (i.e., dh = 25 mm) and
theoretical exit velocities assuming incompressibility (Fig. 3). This assumption is reasonable as a Mach
number of only Ma = 0.11 is estimated for the highest supply rate. It should be noted, that all test cases yield

20

15
f [Hz]

10

0
cout [m/s] 0 5 10 15 20 25 30
ṁ[kg/h] 0 20 40 60 80
Re 0 10000 20000 30000 40000 50000

Figure 3. The jet’s oscillation frequency vs. supply rate for the examined fluidic oscillator.

4 of 15

American Institute of Aeronautics and Astronautics


a Reynolds number well within the turbulent regime of a pipe flow. The oscillation frequencies are obtained
from the internal pressure measurements and reveal the expected linear dependency (Fig. 3). However, a
shift in slope is clearly noticeable. This observation is accompanying by changing jet properties, which are
discussed in subsequent sections.

A. Internal Dynamics
The understanding and knowledge of the fluidic oscillator’s internal dynamics are crucial for designing and
optimizing fluidic oscillators adapted to specific requirements and applications. In the following sections,
the general flow field is discussed and the switching mechanism explained. Furthermore, geometric features
which may hold potential for optimizing and influencing the jet’s properties are suggested. Additionally, the
massflow through the feedback channel (FBC) is discussed in detail as well as the oscillating jet properties at
the outlet.

1. General Flow Field


Before providing a detailed description of the internal mechanisms, a conceptual overview of the main
dynamics is given. The switching mechanism of the fluidic oscillator is based on the Coanda effect. The
jet enters the mixing chamber and entrains surrounding air, hereby generating a low pressure region and
leading to a movement of the jet towards one of the two walls. Without the feedback channels (FBCs), the
jet would remain attached to this wall in a stable state. By adding the FBCs, fluid is drawn from a region
near the outlet nozzle back to the inlet where the jet is attached. The jet interacts with the fluid that is
guided through the FBC and separates to move over to the other wall, reinitiating the process. This switching
motion leads to a spatially oscillating jet at the outlet.
The switching mechanism and the detailed internal dynamics are discussed for an outlet velocity of 11 m/s
with the aid of the depicted flow fields and streamlines in Fig. 4. Half an oscillation period (i.e., the jet
switches from one side to the other) is represented in phase angle steps of φ = 45◦ . The switching mechanism
starts at φ = 0◦ (Fig. 4(a)) when the jet enters the mixing chamber with a positive deflection angle and
attaches to the upper wall. Recall that the criterion for the oscillation start is the zero pressure difference
between the upper and lower FBC inlet. Due to the sharp corner of the inlet wedge, the flow separates and a
recirculation bubble is enclosed downstream of the upper inlet wedge while the recirculation bubble at the
lower wall is at its maximum size. The jet is bent downward and still impinges partially onto the lower outlet
nozzle wall exiting the fluidic oscillator approximately horizontal. As the upper recirculation bubble starts
growing, it pushes the jet downward. A large recirculation zone at the lower wall is remnant from the previous
oscillation cycle. It is pushed downstream, reaching the lower FBC inlet at around φ = 30◦ to φ = 45◦ and
is pushed into the FBC at around φ = 90◦ (Fig. 4(c)). With the growing separation bubble at the upper
wall, the jet now impinges onto the upper wall of the outlet nozzle. While most of the jet exits the fluidic
oscillator with a negative jet deflection angle, a smaller part of the jet enters the oscillator’s upper FBC
inlet and is transported to the FBC outlet in upstream direction. Here, the flow enters the mixing chamber
between the main jet and the upper inlet wedge with a relatively low momentum when compared to the main
jet. Hence, it is fed into the upper recirculation bubble which grows, expands, and pushes the jet away from
the upper wall again. The massflow through the feedback channel is directly related to the growth of the
recirculation bubble, which is discussed in the subsequent section. The expanding recirculation bubble bends
the jet, thereby changing the impingement angle onto the upper outlet nozzle wall. This impingement angle
affects the amount of fluid entering the FBC and thus, the massflow through the FBC. From around φ = 45◦
to 135◦ (Fig. 4(b)-(d)), the impingement angle relative to the upper outlet nozzle wall increases, resulting
in an increasing FBC massflow, which has its maximum between φ = 135◦ and φ = 180◦ (Fig. 4(c),(d)).
Consequently, the recirculation bubble continues growing and pushes the jet over to the lower wall. While
the massflow reaches its maximum in the upper FBC, it is at its minimum in the lower feedback channel. At
around φ = 180◦ (Fig. 4(e)), the jet attaches to the lower wall and a small recirculation bubble is formed
downstream of the lower inlet wedge. This completes one switching process and reinitiates a new one as the
bubble starts growing. The jet changes its direction inside the mixing chamber and is now moving upward
again.

5 of 15

American Institute of Aeronautics and Astronautics


(a)
φ = 0◦ φ = 0◦

(b)
φ = 45◦ φ = 45◦

c/coutlet
1.8

1.6

1.4
(c)
1.2 φ = 90◦ φ = 90◦

0.8

0.6

0.4
(d)
0.2 φ = 135◦ φ = 135◦

(e)
φ = 180◦ φ = 180◦

-10 -8 -6 -4 -2 0 -10 -8 -6 -4 -2 0
x/dh x/dh

Figure 4. The oscillators internal flow field (left) and corresponding streamlines (right) for coutlet = 11 m/s.

6 of 15

American Institute of Aeronautics and Astronautics


Besides the main switching mechanism, some other interesting dynamics are observed. A negative massflow
(i.e., from FBC outlet to inlet) is evident through the upper FBC from φ = 280◦ to φ = 20◦ (Fig. 4(a)) and
the lower FBC from φ = 100◦ to φ = 200◦ (Fig. 4 (d),(e)). As the jet is pushed over to the other wall, a
small portion is cut off by the inlet wedge and penetrates into the FBC in the opposite direction. This flow
through the FBC in downstream direction hinders the jet to fully attach near the FBC outlet. This dynamic
causes two vortices to form in the respective left corner of the FBC as a result of an adverse pressure gradient
and the FBCs’ rectangular shape. These vortices persist until the flow entering the FBC inlet dominates.
The rectangular shape of the FBC leads to the formation of some other recirculating structures. At around
φ = 90◦ (Fig. 4(c)), the jet attaches to the wall near the FBC inlet and the stream separates from the trailing
edge of the geometry. A vortex forms within the cavity of the FBC inlet. Additionally, the sharp edges inside
the FBC cause flow separation, especially when the stream through the FBC reaches its maximum (e.g., FBC
outlet in Fig. 4(c) and FBC inlet in Fig. 4(d)).
It is obvious that the inner geometry of the outlet nozzle influences both the amount of fluid entering the
feedback channel and the jet deflection angle. While the jet impinges onto the inner outlet nozzle wall, the
outer jet deflection angle is almost constant, which implies that the jet’s deflection angle is directly dependent
on the inner wall angle. It is noteworthy that the emitted jet never attaches to the diverging walls of the
outlet nozzle.
The investigation of the switching mechanism exhibits potential for optimization and possibilities to influence
the oscillation properties. Some suggestions include:
1. The rectangular shape of the FBC leads to separations and the formation of vortices, which results
in pressure losses. These losses influence the massflow through the FBC, which is crucial for the
switching process as discussed in the subsequent section. An obstructed mass transport slows the
switching process. Hence, it is suggested that streamlined FBCs may yield a higher oscillation frequency.
Furthermore, the overall oscillator efficiency may be enhanced by reducing the pressure losses within
the device.
2. As previously mentioned, a small portion of the fluid enters the FBC outlet leading to a flow in
downstream direction. This hinders the attachment of the jet to the wall and adds a resistance to the
flow entering the FBC inlet. By increasing the distance of the inlet wedges, the downstream directed
flow through the FBC can be reduced, which is likely to result in a higher oscillation frequency.
3. It is obvious that the outlet nozzle’s geometry highly affects the jet’s deflection angle in the external
flow field. Thus, the outer deflection angle and consequently, the affected area, may be altered by
changing the geometry of the outlet nozzle. This geometry also affects the amount of fluid entering the
FBC inlet which influences the oscillation frequency. Additionally, the impingement of the jet on the
outlet nozzle’s wall leads to an oscillating pressure and substantial pressure loss.
4. Since the jet does not attach to the diverging walls of the outlet nozzle, it is suggested that this part
of the oscillator may not influence the switching behavior at all and may therefore be omitted. This
alteration may simplify the oscillator’s design and system integration.
Many of these suggestions were examined in the numerical study of Bobusch et al.15 These potential changes in
geometry provide the tools for developing an optimized fluidic oscillator for application specific requirements.

2. Massflow in Feedback Channel


As previously discussed, the massflow through the feedback channel is highly important for the switching
mechanism. Bobusch et al.12, 15 suggested that if the governing mechanism of the switching process is the
expanding recirculation bubble, the mass transported through the feedback channels is constant for one
oscillation cycle independent of the supply rate (at least within the incompressible regime). This means
that the recirculation bubble has to grow by a design-specific volume to push the jet to the other side. The
massflow of the lower feedback channel which is calculated at the middle of the duct is shown in the left
of Fig. 5. It is obvious that the peak values in massflow increase with supply rate. This also includes the
reversed flow inside the FBC. Since the cycle duration decreases with the supply rate, a required mass has to
be transported within a shorter time, resulting in the higher peak values. When normalizing the massflow
through the feedback channel, ṁ, by the oscillation frequency, f , the resulting curves coincide (Fig. 5, right).
This suggests that the massflow transported through the FBC per oscillation cycle is independent of the

7 of 15

American Institute of Aeronautics and Astronautics


6 0.4
7 m/s
19 m/s
4 0.3 34 m/s

ṁ/f [g]
ṁ [g/s]
0.2
2
0.1
0
0

-2 -0.1
0 60 120 180 240 300 360 0 60 120 180 240 300 360
φ[◦ ] φ[◦ ]
Figure 5. Left: Massflow through lower feedback channel for three different supply rates; right: the same
massflow normalized by the corresponding oscillation frequency.

supply rate. This statement is verified by integrating the massflow over one oscillation period for all supply
rates. Figure 6 reveals that the total mass of transported air per cycle is generally constant for all supply
rates. Therefore, it can be deduced that the volumetric growth of the recirculation bubble is the governing
mechanism for the switching process and that the oscillation frequency mainly depends on the time it takes
to transport the required mass (or volume) through the feedback channels.
0.2
0.18
total mass per cycle [g]

0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
0 5 10 15 20 25 30 35 40
coutlet [m/s]

Figure 6. Total mass transported through feedback channel per cycle.

3. Oscillation of Jet Properties


The jet’s switching motion is accompanied by an oscillation of the jet properties, as presented in Fig. 7.
The axial pumping is a result of an oscillating overall pressure drop inside the oscillator, which was already
detected by Bobusch et al.12 However, the dynamics and causes of the axial pumping are not fully described
yet. Besides the overall pressure drop, the velocities, massflow, and momentum are oscillating throughout the
oscillator as well. Two main reasons are suggested for this behavior:

1. The changing jet deflection angle leads to a varying effective outlet area and hence changing outlet
velocities. These velocity fluctuations imply massflow and momentum changes.
2. The impingement angle of the jet onto the outlet wall changes. This angle regulates the amount of fluid
entering the feedback channel inlet and exiting the oscillator. Furthermore, an orthogonal impingement
angle leads to a stagnation of the flow and increase of pressure. This also results in fluctuations of the
jet properties.
It is likely to be a combination of both. Note that the smallest jet velocity coincides with the largest jet
width and the largest jet deflection angle coincides with the smallest jet width (Fig. 7). These findings

8 of 15

American Institute of Aeronautics and Astronautics


are consistent with the previous discussion of the internal dynamics. The fluctuation in jet properties by
approximately 10% at the outlet may also affect the understanding of these devices as flow control actuators.
Furthermore, all previous application related studies considered the oscillator’s output (e.g., the momentum
coefficient cµ ) as constant which may not always be appropriate.

1.15
normalized values X/X

cmax
massflow
0.85 momentum
0 60 120 180 240 300 360 jet width
1 40 deflection angle

θjet [◦ ]
w/dh

0.8 0

0.6 -40
0 60 120 180 240 300 360
φ[◦ ]

Figure 7. Jet properties at the outlet.

B. External Dynamics
The internal dynamics directly influence the jet’s properties in the external flow field, which are discussed
in this chapter. First, the general flow field is discussed regarding sweeping behavior and flow phenomena.
Thereafter, the flow field properties (e.g., massflow, jet width, and momentum) are determined and compared
to a non-oscillating jet.

1. General Flow Field


At first glance, the time-resolved external flow field (Fig. 8) indicates that the emitted jet of the fluidic
oscillator is similar to a steady jet when it has its maximum deflection angle and is distorted while flipping to
the other side. The affected region of the jet is substantially larger than the affected region of a conventional,
axisymmetric free jet because of the sweeping motion. At a distance of 10 nozzle diameters from the exit, the
average width of the affected region is 20 nozzle diameters. Therefore, the jet induces energy in a larger area,
which may make it superior to discrete steady jets in flow control and mixing applications. In order to affect
the same region as a non-oscillating jet, the number of nozzles with sweeping jets can thus be reduced.
As the jet sweeps, it distributes vorticity along the affected region (not shown). It is suspected that this
normal vorticity is the source for the effectiveness of the fluidic oscillator in separation control applications as
the normal vorticity is bent to streamwise vorticity by the external flow. A discrete vortex forms on either
side of the jet’s fully deflected state (Fig. 9) due to the present shear layer. The vortex moves downstream
as long as the jet stays fully deflected. As soon as the jet leaves this state in order to flip to the other
side, the discrete vortex disappears. This pair of counter-rotating vortices (Fig. 9, left) was already evident
in other studies7, 18 where flow visualization techniques were applied on an adjacent surface. Due to the
time-averaging effect of the surface flow visualization technique, two seemingly stationary vortices adjacent to
the jet’s sweeping range were depicted. However, due to the steep velocity gradients, vorticity is distributed
even during the sweeping motion although a discrete vortex does not form.

9 of 15

American Institute of Aeronautics and Astronautics


φ = 0◦ φ = 60◦

c
coutlet
1.5

φ = 120◦ φ = 180◦
0.5

0 5 10 15 0 5 10 15
x/dh x/dh

Figure 8. Half oscillation period of the external flow field (coutlet = 7 m/s).

coutlet = 7 m/s (φ = 159◦ )


10 c/coutlet
8 1.5

-2 6

-3 1.0
-4

-5
y/dh

0.5
-6

-7 0
5 10 15 20
-8
x/dh
-9
2 3 4 5 6 7 8
x/dh

Figure 9. Vortex in the external flow field in comparison to a flow visualization.7

10 of 15

American Institute of Aeronautics and Astronautics


The sweeping pattern of the jet is best quantified by its deflection angle. In this study, the instantaneous
deflection angle is defined by the direction of the vector with the maximum velocity magnitude for various
distances from the nozzle. The jet’s maximum deflection angle is presented as a function of supply rate
for various downstream distances (Fig. 10). For low supply rates the deflection angle is almost constant.
However, it decreases significantly for supply rates larger than coutlet = 23 m/s. This trend is visible for
all distances from the nozzle, including the nozzle’s smallest diameter. This observation is also consistent
with the changes in oscillation frequency (Fig. 3). The reasons for that behavior are not fully understood
yet. However, it is assumed that the maximum deflection angle is affected by changes in the internal bubble
dynamics. It is also noteworthy that the maximum deflection increases with distance from the nozzle which
implies a bending of the jet. This behavior is likely caused by entrainment of the surrounding air.

40
θmax [◦ ]

30
r / dh = 0
r / dh = 1.5
20
r / dh = 4
0 5 10 15 20 25 30 35 40
coutlet [m/s]

Figure 10. Maximum deflection angle as a function of the supply rate.

The decrease in deflection angle was initially suspected to be caused by increasing noise levels and decreasing
accuracy of the phase-averaging method. However, examining the transient data by taking each individual
PIV snapshot into account reveals the same result (Fig. 11), which also provides additional confidence in the
applied phase-averaging method. For every point in the external flow field, the occurrences are counted when
the local velocity is higher than 50% of the theoretical outlet velocity. The results reveal the jet’s length
of stay at various positions in the external flow field (Fig. 11). It is evident that the maximum deflection
angle is substantially decreased for higher supply rates. Therefore, the jet appears to have its most impact
along the line of symmetry. In contrast, the smaller supply rates yield two distinct patterns off the center
line, which is caused by the larger deflection angle. Furthermore, it indicates that the jet dwells more in its
deflected state than it takes to flip over to the other side.

coutlet = 4 m/s coutlet = 37 m/s


100
5 5
Occurrences [%]

80

60
y/dh

0 0
40

−5 −5 20
◦ ◦
θjet,max = 36 θjet,max = 24
0 5 10 0 5 10
x/dh x/dh

Figure 11. The jet’s length of stay in the external flow field.

11 of 15

American Institute of Aeronautics and Astronautics


The jet’s deflection angle may be used as an indicator for the instantaneous position of the jet. In Fig. 12, the
jet’s deflection angle and its angular velocity are shown. Three phases of the sweeping pattern in the external
flow field are defined. When the jet’s angular velocity reaches its maximum, an overshoot phase starts. This
phase is characterized by a rapid decrease of the angular velocity when the jet reaches its maximum deflection
angle. However, the jet does not stay at this angle. It moves back to a smaller deflection angle where it
remains temporarily. After a short time the jet leaves the deflected state and accelerates toward the other
side (i.e., the acceleration phase). When the jet is in the overshoot and stagnation phase it is defined as fully
deflected. The acceleration phase is considered as the flipping time. According to this definition, the jet
takes approximately 40% of one oscillation period to flip to the other side. Similar to the jet deflection angle
(Fig. 10), this ratio remains constant for supply rates lower than coutlet = 23 m/s (not shown). For higher
supply rates, this ratio decreases, which is expected because the jet travels a smaller angle range. The small
flipping time also implies that the jet may distribute the fluid unevenly during one oscillation cycle.This may
be of interest for applications relying on an even distribution.
Jet deflection angle
40
Flipping
θf lip
θjet [◦ ]

0 Deflected Deflected
−θf lip
Flipping
-40
0 45 90 135 180 225 270 315 360
Angular velocity magnitude
1
Overshoot Overshoot
|∂θjet /∂φ | [ - ]

0.8 Acceleration Acceleration


0.6
Stagnation Stagnation
0.4
0.2
0
0 45 90 135 180 225 270 315 360
φ [◦ ]

Figure 12. Time-resolved deflection angle at x/dh = 1.4 (coutlet = 15 m/s).

2. Properties of the External Flow Field


Determining the jet properties of an oscillating jet is a challenging task. Due to the sweeping motion, the
analysis is based on polar coordinates in order to keep a constant distance r from the nozzle. Furthermore,
the velocity profile along the polar arc is not symmetrical, as it would be for an axisymmetric steady jet
(Fig. 13(a)). When the jet sweeps from one side to the other, its leading shear layer maintains steep gradients.
In contrast, the trailing shear layer is stretched because the jet pulls previously accelerated flow behind it.
This causes the velocity profiles not only to be asymmetric, but also to be wider than the velocity profile of an
axisymmetric jet. In this study, the instantaneous jet width is defined at 50% of the local maximum velocity
at a prescribed radial distance. Furthermore, the jet properties such as maximum velocity oscillate during the
sweeping motion (Fig. 13(b)) which may also affect the comparison to a steady jet. However, a comparison to
time-averaged properties is considered. Here, ’time-averaged’ refers to the average of a particular parameter
value (e.g., jet width) that was determined at each individual phase angle. In Fig. 14 the time-averaged
maximum velocity and jet width are depicted. It is noticeable that the oscillating jet’s maximum velocity
decreases faster than the maximum velocity of an axisymmetric jet19 (Fig. 14(a)). This indicates a higher
energy transfer to the environment, which is also supported by the jet width (Fig. 14(b)). The jet width of an
axisymmetric jet increases significantly slower than the jet width of an oscillating jet. Furthermore, the slope
increases with distance from the nozzle and thereby deviates substantially from the constant increase of the

12 of 15

American Institute of Aeronautics and Astronautics


axisymmetric jet width. Both, the jet width and the maximum velocity of the oscillating jet are indicative
for a higher entrainment.

14
8 r/dh = 2 1.4
r/dh = 6

Jet width / dh
6 r/dh = 10 1.2

cmax [m/s]
c [m/s]

1 12
4
0.8
2 jet width
0.6 cmax
0 10
−60 −40 −20 0 20 40 60 0 60 120 180 240 300 360
polar angle ψ [◦ ] φ [◦ ]
(a) Instantaneous velocity plots along the polar angle. (b) Oscillating jet properties in the external flow field.

Figure 13. Jet properties in polar coordinates (coutlet = 7 m/s).

8
oscillating jet oscillating jet
axisymmetric jet19 axisymmetric jet19
1 6
cmax /coutlet

w/dh

4
0.5
2

0 0
0 5 10 15 20 0 2 4 6 8 10
r/dh r/dh
(a) The jet’s maximum velocity. (b) The jet width.

Figure 14. Time-averaged jet properties along the distance to the nozzle (coutlet = 7 m/s).

Since a two-dimensional optical measurement system is used, no information regarding the jet depth is
available, making it challenging to determine the total massflow and the entrainment of the oscillating jet.
Therefore, the conservation of momentum is used to calculate an effective jet depth. The time-averaged
momentum per unit depth is not constant because the spreading normal to the sweeping plane is not
considered (Fig. 15(a)). However, by applying the conservation of momentum, a factor is calculated, which
scales the momentum to be constant and equal to the theoretical momentum at the outlet nozzle (Eq. 1).
This factor represents an effective jet depth.
theoretical momentum
effective jet depth z ∗ (r) = (1)
time-averaged calculated momentum(r)

The results of this calculation are plotted in Fig. 15(b). The effective jet depth is independent of the supply
rate and meets the boundary conditions at the nozzle (i.e., z ∗ = dh ). Furthermore, the effective jet depth
presents a lower limit when using the theoretical momentum. The average exit velocity is higher than the
theoretical exit velocity (Fig. 14(a)) due to a smaller effective exit area. Therefore, the proposed jet depth
z ∗ would actually be larger, which in turn would yield a larger total massflow as well. Thus, it is deemed
reasonable to use the effective jet depth for a lower limit calculation of the massflow. In Fig. 15(c) the
massflow, in comparison with a theoretical axisymmetric jet,19 is demonstrated. It is noticeable that the

13 of 15

American Institute of Aeronautics and Astronautics


entrainment rate of the oscillating jet is higher than that of an axisymmetric jet. This is suspected to be
caused by a larger area of contact with the surrounding fluid which provides a larger volume of fluid to be
entrained. Note that this behavior is independent of the supply rate (not shown). The massflow close to
the nozzle is initially smaller because the boundary conditions are not taken into account for the massflow
calculation of the axisymmetric jet. The enclosed nozzle of the fluidic oscillator hinders initial entrainment.
However, the average total mass flow of the oscillating jet is eventually larger than that of an axisymmetric
jet. In order to provide a proper comparison, future studies will include the measurements of a non-oscillating
jet originating from the setup.
·10−3
2 6 15
7 m/s oscillating jet
1.5 15 m/s axisymmetric jet19
J / z [N/m]

ṁ/ṁsupply
4 30 m/s 10

z ∗ /dh
1
2 5
0.5

0 0 0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
r/dh r/dh r/dh
(a) The momentum per unit depth (b) The effective jet depth. (c) The total massflow (coutlet = 7 m/s).
(coutlet = 7 m/s).

Figure 15. Time-averaged jet properties along the radial distance to the nozzle’s center.

V. Conclusion
The present study delivers a detailed description of the internal dynamics and the switching mechanism of
a fluidic oscillator. The volumetric growth of the recirculation bubble is identified as the governing mechanism
and the oscillation frequency is associated with the feedback channel massflow. Several suggestions for
optimization and possibilities to influence the jet’s properties are presented. As the efficiency is positively
influenced by reduced internal losses, it may be enhanced by streamlining the feedback channels. Furthermore,
a lower pressure drop also increases the oscillation frequency, which mainly depends on the time it takes the
fluid to be transported through the feedback channel. Additionally, the oscillation frequency can be raised
by increasing the distance between the inlet wedges. The deflection angle of the jet may be adjusted by
changing the inner geometry of the outlet nozzle. Since the jet does not adhere to the diverging walls of the
outlet nozzle, they may be omitted for this particular design. The investigated internal dynamics result in
oscillating jet properties which are explained as a result of a varying effective outlet area. Furthermore, a
connection between the fluctuation of the overall pressure drop and the impingement angle of the jet onto
the outlet nozzle wall is established.
The oscillating jet properties influence the external flow field as well. Here, the oscillation of the jet is
not harmonic because the jet dwells longer in its deflected state than it needs to switch to the other side.
Furthermore, an overshoot and stagnation of the jet’s motion is detected. This leads to an inhomogeneous
emission of fluid into the affected area which may be of interest or concern for some applications. Below
a threshold supply rate the jet shows a constant behavior. At higher supply rates the jet deflection angle
decreases as a result of changes in the internal dynamics which are not fully understood yet. This change is
also observed in the oscillation frequency which deviates from its linear trend.
The oscillating jet affects a significantly larger area than a steady axisymmetric jet. Furthermore, the
instantaneous velocity profiles reveal a faster decrease in jet velocity accompanied by a larger jet width. The
velocity profiles are asymmetric with steep gradients in its moving direction and shallow gradients on its
trailing side. Consistent with these observations, the analysis of the total mass flow reveals that the oscillating
jet entrains more fluid than a comparable steady jet.

14 of 15

American Institute of Aeronautics and Astronautics


References
1 Stouffer,
R. D., “Oscillating spray device,” Patent US 4151955, 1979.
2 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.
3 Seele, R., Tewes, P., Woszidlo, R., McVeigh, M. A., Lucas, N. J., and Wygnanski, I. J., “Discrete Sweeping Jets as Tools for

Improving the Performance of the V-22,” AIAA Journal of Aircraft, Vol. 46, No. 6, 2009, pp. 2098–2106. doi:10.2514/1.43663.
4 Woszidlo, R., Nawroth, H., Raghu, S., and Wygnanski, I. J., “Parametric Study of Sweeping Jet Actuators for Separation

Control,” AIAA 5th Flow Control Conference (28 June - 01 July 2010, Chicago, Illinois, USA), 2010. doi:10.2514/6.2010-4247.
5 Seele, R., Graff, E., Lin, J., and Wygnanski, I. J., “Performance Enhancement of a Vertical Tail Model with Sweeping Jet

Actuators,” AIAA 51st Aerospace Sciences Meeting (7-10 January 2013, Grapevine, Texas, USA), 2013. doi:10.2514/6.2013-411.
6 Seifert, A., Stalnov, O., Sperber, D., Arwatz, G., Palei, V., David, S., Dayan, I., and Fono, I., “Large Trucks Drag Reduction

Using Active Flow Control,” AIAA 46th Aerospace Sciences Meeting and Exhibit (7-10 January 2008, Reno, Nevada, USA),
2008. doi:10.2514/6.2008-743.
7 Woszidlo, R., Stumper, T., Nayeri, C. N., and Paschereit, C. O., “Experimental Study on Bluff Body Drag Reduction

with Fluidic Oscillators,” AIAA 52nd Aerospace Sciences Meeting (13 - 17 January 2014, National Harbor, Maryland, USA)
(accepted for publication), 2014.
8 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 (9-22 June 2000, Denver, USA), 2000. doi:10.2514/6.2000-2554.
9 Raman, G. and Raghu, S., “Cavity Resonance Suppression Using Miniature Fluidic Oscillators,” AIAA Journal, Vol. 42,

No. 12, 2004, pp. 2608–2612. doi:10.2514/1.521.


10 Guyot, D., Paschereit, C. O., and Raghu, S., “Active Combustion Control Using a Fluidic Oscillator for Asymmetric Fuel

Flow Modulation,” International Journal of Flow Control, Vol. 1, No. 2, 2009, pp. 155–166. doi:10.1260/175682509788913335.
11 Koso, T., Kawaguchi, S., Hojo, M., and Hayami, H., “Flow Mechanism of a Self-Induced Oscillating Jet Issued from a Flip-Flop

Jet Nozzle,” The Fifth JSME-KSME Fluids Engineering Conference (17-21 November 2002, Nagoya, Japan), Vol. 5, 2002.
12 Bobusch, B. C., Woszidlo, R., Bergada, J. M., Nayeri, C. N., and Paschereit, C. O., “Experimental Study of the Internal Flow

Structures inside a Fluidic Oscillator,” Experiments in Fluids, Vol. 54, No. 6, 2013. doi:10.1007/s00348-013-1559-6.
13 Wassermann, F., Hecker, D., Jung, B., Markl, M., Seifert, A., and Grundmann, S., “Phase-locked 3D3C-MRV measurements

in a bi-stable fluidic oscillator,” Experiments in Fluids, Vol. 54, No. 3, 2013. doi:10.1007/s00348-013-1487-5.
14 Koklu, M. and Melton, L. P., “Sweeping Jet Actuator in a Quiescent Environment,” AIAA 43rd Fluid Dynamics Conference

(24-27 June 2013, San Diego, California, USA), 2013. doi:10.2514/6.2013-2477.


15 Bobusch, B. C., Woszidlo, R., Krüger, O., and Paschereit, C. O., “Numerical Investigations on Geometric Parameters Affecting

the Oscillation Properties of a Fluidic Oscillator,” AIAA 21st Computational Fluid Dynamics Conference (24-27 June 2013,
San Diego, California, USA), 2013. doi:10.2514/6.2013-2709.
16 Ostermann, F., Woszidlo, R., Gärtlein, S., Nayeri, C. N., and Paschereit, C. O., “Phase-Averaging Methods for a Naturally

Oscillating Flow Field,” AIAA 52nd Aerospace Sciences Meeting (13 - 17 January 2014, National Harbor, Maryland, USA)
(accepted for publication), 2014.
17 Bergh, H. and Tijdeman, H., “Theoretical and experimental results for the dynamic response of pressure measuring systems,”

National Aerospace Laboratory NLR, Vol. 238, 1965.


18 Woszidlo, R. and Wygnanski, I. J., “Parameters Governing Separation Control with Sweeping Jet Actuators,” AIAA 29th

Applied Aerodynamics Conference (27 - 30 June 2011 ,Honolulu, Hawaii, USA), 2011. doi:10.2514/6.2011-3172.
19 Schlichting, H. and Gersten, K., Grenzschicht-Theorie (German Edition), Springer, 2006.

15 of 15

American Institute of Aeronautics and Astronautics

View publication stats

You might also like