You are on page 1of 12

Active Control of Rectangular Supersonic Twin

Jets using Perturbations: Effects and


Mechanism
N. Webb *, A. Esfahani †, R. Leahy ‡, and M. Samimy §
Gas Dynamics and Turbulence Laboratory, Aerospace Research Center
The Ohio State University, Columbus, Ohio 43235
Rectangular twin jets are quite promising for tactical aircraft. However, the associated screech and
coupling phenomena which can produce elevated far-field noise and strong near-field pressure
fluctuations (with the potential to damage nearby aircraft components) must be mitigated. The
objective of this work is to examine the control authority of localized arc-filament plasma actuators
(LAFPAs) over supersonic rectangular twin jets from closely spaced converging-diverging nozzles
over a wide range of flow regimes, and to explore the underlying physics of this control technique
using perturbations. LAFPAs exert control authority by leveraging flow physics via manipulating
the Kelvin-Helmholtz instability to produce significant effects with minimal power input. LAFPAs
have previously demonstrated excellent control authority in subsonic and supersonic jets. Spectral
proper-orthogonal decomposition of time-resolved schlieren images demonstrate that the twin jets
readily respond to excitation by the LAFPAs over a wide range of frequencies, allowing the LAFPAs
to control the generation and development of the large-scale structures in the jet’s shear layer and
thus their interactions with the shock cells. Time-averaged wavelet coherence magnitude and phase
from near-field acoustic data were used to assess the twin jets’ screech and coupling states. The effects
of various actuation parameters on the twin jets’ screech and coupling are consistent with empirical
predictions based on the classical screech closure model. The LAFPAs’ capabilities include altering
the coupling and the screech modes, changing the screech and coupling frequency, and suppressing
screech and coupling.

I. Nomenclature
De = area-based equivalent nozzle exit diameter
fe = excitation frequency
LSS = large-scale structures
Mj = fully expanded jet Mach number
NPR = nozzle pressure ratio
SPOD = spectral proper orthogonal decomposition
Sts = screech Strouhal number, fsDe/uj
Ste = excitation Strouhal number, feDe/uj
Stm = SPOD mode Strouhal number, fmodeDe/uj
𝛾−1
uj = fully expanded jet velocity, 𝑀𝑗 √𝛾𝑅𝑇𝑜 /(1 + 𝑀𝑗2 )
2

*
Research Scientist, Aerospace Research Center, AIAA Senior Member, webb.356@osu.edu

Ph.D. candidate, Aerospace Research Center, AIAA Student Member

Graduate student, Aerospace Research Center, AIAA Student Member
§
Academy Professor, Aerospace Research Center, AIAA Fellow, samimy.1@osu.edu

1
American Institute of Aeronautics and Astronautics
Φc = relative phase between large-scale structures in adjacent jets
Φsl = relative phase between large-scale structures in the top and bottom shear layers of each jet

II. Introduction
Rectangular twin jets can provide significant benefits to tactical aircraft. The close spacing of the twin jets allows for
compact airframes, able to endure the loading of high-performance flight and the twin engines design provides
important redundancy for safety and survivability considerations. Furthermore, rectangular jets can offer lower
observability, simpler mechanical options for thrust vectoring, and reduced drag relative to axisymmetric nozzles
[1,2]. However, for supersonic jets, especially those with sharp throats (i.e., presence of shocks even at design
conditions), jet screech, a strong tonal noise, is an important consideration. When two jets are located near one another,
their screech loops can couple, with the potential to further increase the far-field radiated noise and near-field pressure
fluctuations [3,4]. Coupled twin jets can produce near-field pressure fluctuations which are strong enough to damage
nearby aircraft components (e.g., the B-1A and the F-15E) [5]. While maintaining sufficient distance between the jets
can minimize their interaction, small spacing is desirable from a system design perspective. It is therefore important
to understand and mitigate the adverse consequences associated with screech and coupling.
To effectively implement active flow control of any kind it is imperative to have a good understanding of the
underlying flow physics. Therefore, it is imperative to have a good understanding of jet screech and the directly related
coupling phenomenon. Powell [6] first proposed a feedback closure model for screech. Subsequently, various
expansions/modifications to his original model have been proposed [7,8] but none have substantially changed the
original feedback loop concept. Specifically, screech occurs when perturbations in the jet shear layer near the nozzle
exit excite the Kelvin-Helmholtz (K-H) instability and generate large-scale structures (LSS) in the shear layer of the
jets. The LSS convect downstream, interacting with the jet shock system. These interactions produce acoustic waves
which propagate upstream and perturb the jet shear layer, generating more LSS and thereby completing the feedback
loop. To be self-sustaining, this feedback process must occur over an integer multiple of the screech frequency period.
Powell’s own original model, as well as other subsequent works, have held that the feedback waves are acoustic in
nature and propagate upstream outside of the jet [6–8]. More recent work has discovered trapped or guided waves,
also termed neutral modes, within the jet which could complete the screech feedback loop [9–13]. No matter which
type of wave is confirmed as the dominant source of feedback in the screech loop, the role of the LSS and the shock
system are clear. Additionally, both types of waves are known to propagate upstream at (or near, in the case of guided
waves) the ambient speed of sound [12–14]. Twin jets coupling is also connected to the screech feedback. When the
screech frequency and nozzle spacing are such that feedback waves from one jet arrive at its own nozzle exit in-phase
with the acoustic waves from the other jet, these perturbations reinforce one another and further strengthen the
feedback loop. This couples the screech loops and can significantly increase the near-field pressure fluctuations (along
the minor-axis plane within the inter-nozzle region) as well as the far-field noise. Note that external acoustic waves
are responsible for coupling, as the trapped waves/neutral modes are contained within the jet and cannot not facilitate
the coupling process.
Localized arc-filament plasma actuators (LAFPAs) were developed at the Gas Dynamics and Turbulence Laboratory
as a method of controlling the organization and timing of LSS forming in shear flows [15,16]. They have demonstrated
significant control authority in a wide variety of flows: subsonic and supersonic jets at various flow conditions and
configurations [17–20] and subsonic and supersonic cavity flows [21,22]. They function by periodically injecting
thermal energy into the flow via a low-power arc. These thermal perturbations excite the K-H instability, causing the
generation of LSS in the shear layer of the jets. By using the LAFPAs in groups and altering the frequency at which
they are operated, the timing and organization of the resulting LSS can be altered. Due to the fundamental physics
governing LSS growth and development, the LAFPAs can control the structures’ coherence, entrainment capabilities,
and persistence (i.e., disintegration location). It is important to note that the LAFPAs achieve significant control
authority using only minimal power input by leveraging the naturally present flow physics. This strategy allows the
flow physics to be better understood by observing the LAFPAs’ effectiveness, or lack thereof, at various operating
conditions. As discussed above, LSS play a crucial role in the screech (and therefore coupling) feedback process and
are responsible for the extraction of energy from the mean flow to drive the screech feedback loop [23]. Thus, the
LAFPAs’ ability to alter LSS timing, organization, coherence, etc. allows them to effectively control the screech and
coupling processes. Additionally, LSS are also responsible for the peak mixing noise at shallow polar angles
(measured from the downstream jet axis). Thus, by controlling LSS, the LAFPAs can also mitigate peak mixing noise.

2
American Institute of Aeronautics and Astronautics
The LAFPAs have been employed to control rectangular twin jets, expanding upon previously published work
examining their effects [24]. The jets are operated at various nozzle pressure ratios (NPRs) with differing screech
amplitudes and coupling strength and mode [25]. The objective of this work is to demonstrate the control authority
and effects of the LAFPAs on screech and coupling. Additionally, the physics behind the control authority is explored,
to demonstrate that a good understanding of the relevant physics can be used to intelligently guide the application of
flow control.

III. Experimental Arrangement

A. Facility
The control authority of the LAFPAs was tested on rectangular twin jets in an anechoic test facility at Ohio State’s
Gas Dynamics and Turbulence Laboratory within the Aerospace Research Center. The twin jets (shown in Figure 1)
consist of two rectangular jets of aspect ratio 2. The jet width is 0.95 inches (24.1 mm), and the adjacent nozzle lips
are spaced 0.758 inches (19.25 mm) apart: 1 area-based equivalent diameter (𝐷𝑒 = 2√(𝑏 ∗ ℎ)/𝜋), where b and h are
the width and height of the nozzle exit. The jet nozzles are bi-conic with a sharp throat. The design Mach number is
1.5. The twin jets assembly is installed within a 6.2 m by 5.6 m by 3.4 m anechoic chamber. The flow is driven by
high pressure air from two large (36 m3 total capacity) pressure vessels with a maximum pressure of about 2300 psi
(16 MPa). The stagnation pressure of the flow is set by a computer-controlled valve, which can be adjusted to maintain
the desired NPR. For this work, the NPR was varied from 2.97 to 5.32 (Mj = 1.35 to 1.75). This allowed the LAFPAs’
control authority to be examined on cases from the over- to the underexpanded operating regime.

(a) (b)

Figure 1: Experimental arrangement: a) Internal, bi-conical nozzle contours, b) Twin jets, including nozzles, major
and minor axes, and near-field microphone locations.

B. Plasma Actuators
The diverging section of the converging-diverging nozzles is constructed from boron nitride, a ceramic with a high
dielectric constant and thermal properties which enable it to withstand the high-voltage, high-temperature arc
generated by the LAFPAs. Each jet has 6 LAFPAs distributed around the perimeter. Three each in the top and bottom
(parallel to the major axis) lips of the nozzle. Each actuator consists of two, 1-mm diameter, tungsten electrodes, one
of which is grounded, the other is connected to the in-house built high-voltage pulse generator. Each actuator channel
is individually computer controlled, allowing a wide variety of excitation conditions (various frequencies and relative
phases between channels) to be implemented. Various excitation frequencies (Ste = feDe/Uj) were tested. Throughout
this work, the 3 actuators on any given nozzle lip were always operated in-phase with one another. The actuators were

3
American Institute of Aeronautics and Astronautics
operated in 4 different patterns or modes by changing the relative phase between the LAFPAs on each shear layer. To
identify these modes, 2 phases are defined (see Figure 2). The coupling phase (Φc) describes the phase delay between
the left and the right jet. Excitation modes with Φc = 0° will henceforth be termed “in-phase” and those with Φc =
180°, “out-of-phase”. The shear-layer phase (Φsl) describes the phase delay between the shear layers on the top and
bottom lips of a single jet. Throughout the remainder of the paper, excitation modes with Φsl = 0° will be referred to
as “symmetric” and excitation modes with Φsl = 180° will be referred to as “flapping”. The flapping mode corresponds
to the observed natural shear layer and screech mode of the jets in all baseline cases across the tested range of NPRs.
Figure 3 provides two examples of excitation modes employed in this work.

Figure 2: Φc and Φsl allow the precise operating pattern of the actuators to be specified.

Figure 3: Sample excitation modes used to excite the twin jets: In-phase symmetric excitation (left) and out-of-phase
flapping excitation (right). Semi-transparent starbursts signify actuators pulsed 180° out-of-phase from the others.

C. Diagnostics
The effects of the actuators operating in various conditions on the twin jets flow and acoustics were recorded using
two primary flow diagnostics: near-field microphone measurements and time-resolved schlieren imaging. The
schlieren images were collected using a standard Z-type schlieren arrangement. The jets were imaged such that the
collimated light beam passed through them both (i.e., parallel to the jets’ shared major axis, see Figure 1b). This
provided a good view of the LSS along the major axis shear layers and allowed the LAFPAs’ effects on them to be
documented. A HPLS-36 high-powered pulsed LED from Lightspeed Technologies was used in continuous mode for
illumination and the images were acquired at approximately 40,000 frames per second with a 1µs exposure by a
Phantom v1210 high-speed camera. The knife edge was oriented vertically to highlight horizontal density gradients
and 2000 images with a window size of 512×480 pixels were acquired for each test condition. The results were post-
processed on the Ohio Supercomputer Center using a spectral proper orthogonal decomposition (SPOD) code
developed by Schmidt and Colonius [26]. This technique allowed the coherence and pattern (flapping or symmetric)
of the LSS to be deduced.
The second primary technique by which the actuators’ effects were observed was near-field microphone
measurements. Two Brüel and Kjær 4939 ¼ in. condenser microphones were located, one along the minor axis of
each jet, 4De from the nearest jet centerline (see Figure 1b). These microphones provided information about the
upstream-travelling acoustic waves of each jet. The sampling frequency was 200 kHz and 100 blocks of 32,768

4
American Institute of Aeronautics and Astronautics
samples were acquired for each data point, resulting in a frequency resolution of 6.10 Hz. Signals collected by
microphones were amplified and band-pass filtered between 20 Hz and 100 kHz, using a Nexus 2690 signal
conditioner. The Morlet-wavelet coherence between the microphone signals was calculated. Then, the time averaged
coherence and phase was calculated (note coherence and phase were set to zero when the coherence magnitude was
below 0.7). These quantities were used to determine the jet coupling strength and mode (in-phase or out-of-phase).
This allowed the jets’ response to the LAFPA-organized LSS to be observed. Specifically, the strength and mode of
the jets’ coupling under various excitation conditions was determined.

IV. Results

A. Control Effects on Large-Scale Structures


The LAFPAs have demonstrated their control authority over the characteristics of LSS in a variety of flows. It is via
control of the LSS that the LAFPAs exert significant control authority. SPOD mode shapes calculated from time-
resolved schlieren images provide information about the LAFPAs’ effects on the LSS in the currently studied twin
jets. Figures 4 and 5 show the first SPOD mode (at different frequencies) for various baseline and controlled cases at
two Mach numbers. It is important to realize that SPOD extracts coherent fluctuations from the schlieren images, not
necessarily LSS directly. However, as the LSS are highly coherent in these screeching jets, i.e., the energy is
concentrated in the first few modes, the SPOD does provide information about the LSS’ coherence and organization.
In the subsequent discussion, “SPOD mode shapes” and “LSS” will be used interchangeably with the understanding
that the SPOD mode shapes do not directly represent LSS. It is also important to recall that schlieren is a line-of-sight
averaging technique. Therefore, in cases where the jets are coupled out-of-phase, the density gradient is averaged over
the two jets operating in different phases and thus the apparent nature of the of structures may be altered. Therefore,
the results presented below are all from in-phase coupled cases. In a few cases below, the relatively low frequency
resolution of the SPOD analysis (625 Hz) should result in an apparent (slight) mismatch between the mode frequency
and the excitation frequency. In such cases, the mode frequency has been reported as equal to the excitation frequency
to avoid confusion.

(a) Baseline (b) Excited

Figure 4: Magnitude of the most energetic SPOD mode (first mode) at Stm = 0.22 for the Mj = 1.65 twin jets: a)
baseline and b) excited in-phase flapping (Φc = 0°, Φsl = 180°) at Ste = 0.22.
Figure 4a and b compare SPOD modes at the natural screech frequency for the baseline Mj = 1.65 jets with those
excited with an in-phase flapping mode (Φc = 0°, Φsl = 180°) at the natural screech frequency. Note the significant
strengthening of the coherent modes in the excited case. This demonstrates that, for cases in which the natural
screech/coupling loop is relatively weak, appropriate excitation by the LAFPAs can increases the coherence of the
LSS.
Figure 5a and b compare SPOD modes at the natural screech frequency for the baseline Mj = 1.35 jets with those
excited with an in-phase symmetric mode (Φc = 0°, Φsl = 0°), at the natural screech frequency. Across all tested NPR
cases, these twin jets were found to adopt a flapping screech mode, and Figure 5a reflects this. However, when excited
in a symmetric mode, the mode shapes (Figure 5b) clearly reflect LSS which are now organized symmetrically across
the jets’ major axis. Note that these structures (which are vertically stacked rather than staggered as in the natural flow
condition) are quite coherent. Additionally, a brief thought experiment confirms that the timing of the screech feedback

5
American Institute of Aeronautics and Astronautics
loop is not affected by the change from a flapping to symmetric screech mode (i.e., the shock locations, LSS convective
velocity, and acoustic propagation velocity do not depend on the screech mode).

(a) Baseline, Stm = 0.41 (b) In-phase symmetric excitation (Φc = 0°, Φsl = 0°) at
Ste = 0.41, Stm = 0.41

(c) Baseline, Stm = 0.51 (d) In-phase flapping excitation (Φc = 0°, Φsl = 180°) at
Ste = 0.51, Stm = 0.51

(e) In-phase flapping excitation (Φc = 0°, Φsl = 180°) at (f) In-phase flapping excitation (Φc = 0°, Φsl = 180°) at
Ste = 0.97, Stm = 0.41 Ste = 0.97, Stm = 0.97

Figure 5: Magnitude of the most energetic SPOD mode (first mode) for various excitation conditions at Mj = 1.35.
Ste indicates the excitation frequency, while Stm indicates the frequency of the SPOD mode.
In addition to strengthening or altering the organization of the LSS, the LAFPAs can change the passage frequency of
naturally occurring LSS. Figure 5c and d display the SPOD modes at Stm = 0.51 for the baseline Mj = 1.35 jets and
those excited with an in-phase, flapping mode (Φc = 0°, Φsl = 180°) at Ste = 0.51. The baseline flow does not contain
LSS at St = 0.51 (Figure 5c). Rather, the dominant LSS in the baseline case exist at the natural screech frequency
(Figure 5a). Excitation replaces these naturally occurring structures with LSS at the excitation frequency (Figure 5d).
This showcases the LAFPAs ability not only to suppress or enhance naturally existing LSS, but also to replace them

6
American Institute of Aeronautics and Astronautics
with completely different LSS. This is relevant to potential applications as the new structures can have different
coherence, screech/coupling relevant characteristics, acoustic radiation efficiency, etc.
One final comparison serves to highlight the nature of the LAFPAs’ control mechanism. Figure 5e and f compare the
SPOD modes at the natural screech frequency and at the excitation frequency for the Mj = 1.35 case excited with an
in-phase flapping mode (Φc = 0°, Φsl = 180°) at Ste = 0.97. Both modes have clearly defined (spatially) highly
correlated (i.e., dark blue/red) shapes. This indicates that coherent LSS exist at both the natural screech frequency and
the excitation frequency. The jets clearly respond to excitation by generating LSS at the excitation frequency (Figure
5f). However, due to the high-frequency nature of the excitation-produced structures, the natural screech/coupling
loop has not been replaced. This illustrates how the LAFPAs are working with the existing jet physics to organize the
structures. Specifically, the excitation frequency (Ste = 0.97) is high enough to be outside the jet column mode (i.e.,
the jets’ “preferred mode” of the K-H instability) but not yet high enough to be near the jet shear-layer mode (most-
amplified frequency) [15,16,27]. In this situation, while the jet responds to LAFPA excitation by forming coherent
structures, they are not dominant (as when excitation was within the jet column mode, e.g., Figure 5d) and therefore
do not disrupt or replace the natural screech/coupling loop.

B. Twin Jets’ Flow Physics


As demonstrated above, the jets do respond to excitation, demonstrating the LAFPAs’ control authority to dictate the
organization and timing of the LSS. Significant previous work has already demonstrated this control authority [17],
and investigated the control mechanism [16]. The LSS which are controlled by the LAFPAs are an essential
component of the screech and coupling loop. However, as with the control of the structures themselves, the LSS cannot
arbitrarily force, for example, the coupling mode to switch, or to suppress the screech. They must work within the
existing physics of the twin jets to alter the twin jets’ behavior. Thus, to properly understand what the results of
excitation at a given frequency and mode will be, some model of the fundamental physics of twin jets screech and
coupling must be employed. Part of this work has included the development of such a model and a detailed explanation
and validation of the screech closure model will be forthcoming. The developed model is based on the external acoustic
waves and is similar to that of Norum et al. [28] for single jets. This empirical closure model can predict the response
(screech and coupling) of the twin jets to active control using perturbations provided by LAFPAs.
A brief description of the screech closure model is included here. Powell [6] first proposed that the total time required
for a structure/wave to convect/propagate through the screech process must be an integer multiple of the screech period
for the screech loop to be self-sustaining. This postulate will hereafter be referred to as Powell’s timing criterion. An
empirical screech closure model based upon this concept and Norum et al.’s [28] model has been developed.
Consistent with the existence of broadband shock associated noise (BBSAN) [29] it assumes that the LSS interact
with each shock cell as they convect downstream, generating multiple acoustic waves. The generated acoustic waves
then propagate upstream to the nozzle lip, interfering constructively or destructively with one another. Empirical
knowledge of the shock cell locations and convective velocity of the LSS allows this interference pattern, and the
resultant shear layer perturbation amplitudes, to be calculated. Expanding this advancement of Powell’s screech
closure mechanism to twin jets was straightforward: the acoustic waves produced by the second jet are simply added
to the calculation of the feedback interference pattern. As stated above, this screech/coupling closure model is similar
to that developed by Norum et al. [28]. It has been extended in 2 respects: 1) the elimination of Norum et al.’s [28]
assumptions that the shock cell spacing is constant and that the convective velocity does not depend on Mach number,
and 2) the extension of the model to include twin jets coupling.
Figure 6 displays calculated perturbation amplitudes (at the jet nozzle lip) for the twin jets under study, operated at Mj
= 1.5, for two assumed cases: with the jets coupled (i.e., LSS genesis) in-phase and out-of-phase. The vertical dashed
lines mark frequencies which satisfy Powell’s timing criterion and can therefore support self-sustaining feedback
loops. The perturbation amplitude plot provides information about what excitation frequency and pattern LAFPAs
should use to achieve desired results. For example, if the LAFPAs organize the LSS out-of-phase at St = 0.41, the
feedback perturbation amplitude (Figure 6) will be low. This organization of LSS will generate acoustic waves which
destructively interfere with one another, thereby suppressing coupling. However, if the LAFPAs organize structures
in-phase or out-of-phase at St = 0.31, the perturbation amplitude will be high and the acoustic waves will constructively
interfere, resulting in strong screech/coupling. This empirical prediction capability allows the LAFPAs’ operating
parameters to be intelligently selected, as well as illustrating how the LAFPAs work with naturally present physics.
Recall that leveraging flow physics is what allows the LAFPAs to produce such significant changes with very minimal
power input [15,16].

7
American Institute of Aeronautics and Astronautics
Figure 6: Calculated perturbation amplitude at the twin jets’ nozzle lip for Mj = 1.5 over a large range of frequencies
and two coupling modes.

C. Control Effects on Twin Jets’ Coupling


As discussed in Section III-B, though the LAFPAs have clearly demonstrated the ability to control the LSS
organization and timing (see Section III-A), the LSS cannot arbitrarily change the jet screech and coupling condition
but must work within the flow physics of the jets. Thus, it is imperative to examine the response of the jets to the
modified LSS organization. This was accomplished using wavelet coherence and phase results obtained from the near-
field microphone data (see Section II-C for details). Figure 7 compares several such plots for the jets at Mj = 1.5. Each
comparison has been chosen to illustrate both the capabilities of the LAFPAs to significantly modify the twin jets’
coupling and screech, as well as leverage flow physics, based on the screech closure model discussed above. The
baseline, Mj = 1.5 coupling condition is shown in Figure 7a. The baseline flow is strongly coupled out-of-phase. This
can be inferred from the high correlation peak (~ 0.9) at the screech frequency (St = 0.31) and phase at the screech
frequency being near . Spectra (not shown) confirm that the screech peak is strong (> 20 dB) as expected in a strongly
coupled case. Figure 7b shows the jets’ response to out-of-phase flapping (Φc = 180°, Φsl = 180°) excitation at the
natural screech frequency. This excitation reinforces the LSS organization imposed by the natural screech and
coupling modes. As expected, excitation at this condition enhances the coupling strength. The LAFPAs clearly
demonstrate the ability to significantly strengthen coupling.

(a) Baseline, natural screech at Sts = 0.31 (b) Excited out-of-phase, flapping (Φc = 180°, Φsl =
180°) at Ste = 0.31

8
American Institute of Aeronautics and Astronautics
(c) Excited in-phase, flapping (Φc = 0°, Φsl = 180°) at (d) Excited out-of-phase, symmetric (Φc = 180°, Φsl =
Ste = 0.31 0°) at Ste = 0.31

(e) Excited in-phase, flapping (Φc = 0°, Φsl = 180°) at (f) Excited out-of-phase, flapping (Φc = 180°, Φsl =
Ste = 0.36 180°) at Ste = 0.36

Figure 7: Morlet wavelet, time-averaged coherence and phase plots for the Mj = 1.5 baseline (naturally out-of-phase
coupled, Φc = 180°) and various excited cases.
In addition to reinforcing the natural screech/coupling, the LAFPAs are also able to alter the twin jets’ screech and
coupling modes and frequencies. An examination of Figure 7c (in-phase, flapping excitation, Φc = 0°, Φsl = 180°, at
the natural screech frequency) shows that the LAFPAs have altered the natural coupling mode, and recoupled the jets
in-phase even more strongly than when they matched the natural screech and coupling modes (c.f. Figure 7b). This is
possible only in this case, as the screech closure model predicts that the twin jets should respond strongly to both in-
phase and out-of-phase excitation at the natural screech frequency (see Figure 6). However, the in-phase response is
predicted to be stronger. Thus, the observed greater coupling enhancement makes sense. The phase curve confirms
that the LAFPAs have altered the coupling mode of the twin jets to in-phase. The LAFPAs are also able to alter the
natural screech mode of the jets. As has been shown by the SPOD results (Figure 5a and b) the LAFPAs are able to
organize the structures in a symmetric fashion. Figure 7d (out-of-phase, symmetric excitation, Φc = 180°, Φsl = 0°, at
the natural screech frequency) demonstrates that the jets do respond to these structures and couple quite strongly. This
result is interesting given that the jets prefer the flapping screech mode (as confirmed by the adoption of this mode in
all baseline cases [25]). Nevertheless, the LAFPAs have proven able to alter both the screech and coupling mode.
Figure 7e (in-phase, flapping excitation, Φc = 0°, Φsl = 180°, at St = 0.36) demonstrates the LAFPAs’ ability to alter
the screech/coupling frequency as well. Their effect is somewhat obscured by the low frequency resolution inherent
to wavelet analysis, however, a careful comparison of the coherence peak in Figure 7e with the baseline (Figure 7a)
shows that coupling has indeed shifted to the excitation frequency. Spectra of the microphone data (not shown) also
confirm the screech and coupling frequency has been altered to match the excitation frequency.
Finally, comparing Figure 7f (out-of-phase, flapping excitation, Φc = 0°, Φsl = 180°, at St = 0.36) with the baseline
confirms that the LAFPAs’ also possess the ability to significantly suppress coupling. This is, as in the other cases, in
agreement with the predictions of the screech closure model (c.f. Figures 6 and 7). At this frequency, the jets are
predicted to couple poorly out-of-phase, i.e., the acoustic waves produced by the LAFPA-organized LSS interfere
destructively and reduce the strength of the screech and coupling feedback loops. In alignment with this prediction,
the excitation significantly suppresses the coupling of the twin jets. This suppression is associated with reduced near-

9
American Institute of Aeronautics and Astronautics
field pressure fluctuations. Note the distinction which must be made between near-field pressure reduction and far-
field noise suppression. The LAFPAs are significantly suppressing coupling, by organizing the LSS to produce
feedback waves which destructively interfere. This likely decreases the coherence of the LSS, but they are still present.
Thus, while screech and coupling are suppressed, BBSAN and mixing noise are not significantly reduced. This is what
has motivated the next phase of this research to explore excitation at very high frequencies (within the shear-layer
mode frequency band [15,16,27]) to simultaneously suppress near-field pressure fluctuations and reduce far-field
noise.
The actuators’ control authority extends across the range of tested NPR cases. Figure 8 shows two excitation cases
which confirm this, as well as illustrating various effects the LAFPAs can have. First consider the case at Mj = 1.45
(Figure 8a and b) with out-of-phase, flapping (Φc = 180°, Φsl = 180°) excitation at Ste = 0.41 (the natural screech
frequency is St = 0.34). The baseline jets have coupled strongly out-of-phase (Figure 8a). This natural screech/coupling
loop is replaced by a very weak coupling loop at the excitation frequency (Figure 8b). This is clear from the shift in
frequency of the coherence peak. Thus, the overall coupling is suppressed. Even more impressive are the LAFPAs’
effects in the Mj = 1.75 case (Figure 8c and d). The baseline jets have moderately strongly coupled in-phase at this
NPR (Figure 8c). Out-of-phase flapping (Φc = 180°, Φsl = 180°) excitation at Ste = 0.25 (the natural screech frequency
is St = 0.20) suppresses the natural coupling without recoupling the jets at the excitation frequency Figure 8d). This
can be determined conclusively by examining the phase curve, which is nowhere near ± (i.e., out-of-phase coupled)
at the excitation frequency. Figure 8 and Figure 7a and f provide excellent examples of the LAFPAs’ ability to suppress
coupling in supersonic twin rectangular jets.

(a) Mj = 1.45, baseline, natural screech at Sts = 0.34 (b) Mj = 1.45, out-of-phase, flapping excitation (Φc =
180°, Φsl = 180°) at Ste = 0.41

(c) Mj = 1.75, baseline, natural screech at Sts = 0.20 (d) Mj = 1.75, out-of-phase, flapping excitation (Φc =
180°, Φsl = 180°) at Ste = 0.25

Figure 8: LAFPAs’ control authority at various NPRs.

V. Conclusions
This work investigates the control mechanism and authority of localized arc-filament plasma actuators (LAFPAs) over
supersonic rectangular twin jets. Bi-conical nozzles of aspect ratio 2 were operated at a variety of nozzle-pressure
ratios (NPR) from 2.97 to 5.32 (Mj = 1.35 to 1.75). This provided test conditions in both the overexpanded and

10
American Institute of Aeronautics and Astronautics
underexpanded regimes. The flow was interrogated using time-resolved schlieren imaging and two near-field
microphones, one positioned along the minor axis of each jet.
Spectral proper-orthogonal decomposition mode shapes calculated from the schlieren results demonstrated that the
LAFPAs could control the organization of the large-scale structures (LSS) within the jet shear layers by leveraging
the Kelvin-Helmholtz instability. If the LAFPA-generated LSS were within the jet column mode frequency band, the
existing structures in the jet were either reinforced (for excitation at the natural screech frequency) or replaced. If the
introduced structures were at a significantly higher frequency, they were observed to coexist with the naturally present
structures. The LAFPAs also demonstrated the ability to alter the organization of the structures, from flapping (Φsl =
180°, out-of-phase LSS within the top and bottom shear layers) to symmetric (Φsl = 0°, in-phase LSS within the top
and bottom shear layers).
Morlet wavelet coherence and phase, calculated from acquired microphone data, were employed to determine the
effect of the LAFPAs on the twin jets’ coupling. A screech closure model was developed and used to predict the twin
jets’ response to excitation and LAFPAs’ effects based on the underlying flow physics. The observed effects were in
alignment with the predicted jets’ responses over a wide range of jet and excitation conditions. This highlights the
importance of understanding the flow physics for the effective implementation of flow control. The LAFPAs were
demonstrated to have significant control authority over the twin jets’ screech and coupling. They were able to enhance
the coupling, modify the screech or coupling mode (independently), change the screech and coupling frequency, and
significantly suppress coupling. Control authority was observed across a range of NPRs.
While the actuators have demonstrated significant control authority over the twin jets’ screech and coupling, they have
done so by introducing structures at or near the jet column mode frequency band. Thus, there are still coherent LSS
(though with reduced coherency) within the controlled jets. While the present results suggest their ability to
significantly reduce near-field pressure fluctuations (experiments acquiring near-field pressure maps to confirm this
are underway) there is a distinction between near-field pressure fluctuations and far-field noise reduction. Eliminating
in-phase coupling should, according to the literature, reduce the far-field noise. However, if the jets could be decoupled
by replacing the LSS with very small structures which quickly disintegrate, this would produce a much more
significant reduction of the far-field noise. Thus, the next phase of this project will include the introduction of
excitation near the shear layer mode frequency band (much higher frequencies) to allow the decoupling of the jets
without the introduction of coherent LSS. This will result, not only in the decoupling of the jets and suppression of
screech, but also (as the LSS are eliminated) in the mitigation of broadband shock-associated noise and significant
reduction of mixing noise.

Acknowledgements
The support of this work by the Office of Naval Research (Steve Martens) is gratefully acknowledged. The computing
resources for the calculation of the spectral proper-orthogonal decomposition results were provided by the Ohio
Supercomputer Center. The authors wish to thank Oliver Schmidt (University of California, San Diego) for sharing
his SPOD code and Andrew Sais for assistance in processing high-speed schlieren images.

References
[1] Dusa, D., Speir, D., Rowe, R., and Leavitt, L. Advanced Technology Exhaust Nozzle Development. In 19th
Joint Propulsion Conference, No. 1983–1286, 1983.
[2] Wiegand, C., Bullick, B., Catt, J., Hamstra, J., Walker, G., and Wurth, S. F-35 Air Vehicle Technology
Overview. In AIAA Aviation Forum, No. 2018–3368, Atlanta, GA, 2018, pp. 1–28.
[3] Shaw, L. “Twin-Jet Screech Suppression.” Journal of Aircraft, Vol. 27, No. 8, 1990, pp. 708–715.
https://doi.org/10.2514/3.25344.
[4] Panickar, P., Srinivasan, K., and Raman, G. “Aeroacoustic Features of Coupled Twin Jets with Spanwise
Oblique Shock-Cells.” Journal of Sound and Vibration, Vol. 278, 2004, pp. 155–179.
https://doi.org/10.1016/j.jsv.2003.10.011.
[5] Walker, S. Twin Jet Screech Suppression Concepts Tested for 4.7% Axisymmetric and Two-Dimensional
Nozzle Configurations. In AIAA/SAE/ASME/ASEE 26th Joint Propulsion Conference, No. 90–2150, Orlando,
FL, 1990, pp. 1–13.
[6] Powell, A. “On the Mechanism of Choked Jet Noise.” Proceedings of the Physical Society B, Vol. 66, 1953,
pp. 1039–1056.

11
American Institute of Aeronautics and Astronautics
[7] Tam, C., Seiner, J., and Yu, J. “Proposed Relationship between Broadband Shock Associated Noise and
Screech Tones.” Journal of Sound and Vibration, Vol. 110, No. 2, 1986, pp. 309–321.
https://doi.org/10.1016/S0022-460X(86)80212-7.
[8] Panda, J. “An Experimental Investigation of Screech Noise Generation.” Journal of Fluid Mechanics, Vol.
378, 1999, pp. 71–96. https://doi.org/10.1017/S0022112098003383.
[9] Tam, C., and Hu, F. “On the Three Families of Instability Waves of High-Speed Jets.” Journal of Fluid
Mechanics, Vol. 201, 1989, pp. 447–483. https://doi.org/10.1017/S002211208900100X.
[10] Zaman, K., Fagan, A., and Upadhyay, P. “Pressure Fluctuations Due to ‘Trapped Waves’ in the Initial Region
of Compressible Jets.” submitted to Journal of Fluid Mechanics, 2022.
[11] Bogey, C. “Acoustic Tones in the Near-Nozzle Region of Jets: Characteristics and Variations between Mach
Number 0.5 and 2.” Journal of Fluid Mechanics, Vol. 921, No. A3, 2021.
https://doi.org/10.1017/jfm.2021.426.
[12] Edgington-Mitchell, D., Jaunet, V., Jordan, P., Towne, A., Soria, J., and Honnery, D. “Upstream-Travelling
Acoustic Jet Modes as a Closure Mechanism for Screech.” Journal of Fluid Mechanics, Vol. 855, No. R1,
2018, pp. 1–12. https://doi.org/10.1017/jfm.2018.642.
[13] Wu, G., Lele, S., and Jeun, J. Internal and External Feedback in Rectangular Jet Screech. In AIAA Aviation
Forum, No. 2021–2153, Virtual Event, 2021, pp. 1–15.
[14] Mancinelli, M., Jaunet, V., Jordan, P., and Towne, A. “Screech-Tone Prediction Using Upstream-Travelling
Jet Modes.” Experiments in Fluids, Vol. 60, No. 22, 2019, pp. 1–9. https://doi.org/10.1007/s00348-018-2673-
2.
[15] Samimy, M., Kim, J.-H., Kearney-Fischer, M., and Sinha, A. “Acoustic and Flow Fields of an Excited High
Reynolds Number Axisymmetric Supersonic Jet.” Journal of Fluid Mechanics, Vol. 656, 2010, pp. 507–529.
https://doi.org/10.1017/S0022112010001357.
[16] Samimy, M., Webb, N., and Crawley, M. “Excitation of Free Shear-Layer Instabilities for High-Speed Flow
Control.” AIAA Journal, Vol. 56, No. 5, 2018, pp. 1770–1791. https://doi.org/10.2514/1.J056610.
[17] Samimy, M., Kim, J.-H., Kastner, J., Adamovich, I., and Utkin, Y. “Active Control of High-Speed and High-
Reynolds-Number Jets Using Plasma Actuators.” Journal of Fluid Mechanics, Vol. 578, 2007, pp. 305–330.
https://doi.org/10.1017/S0022112007004867.
[18] Kearney-Fischer, M., Kim, J.-H., and Samimy, M. “Noise Control of a High Reynolds Number High Speed
Heated Jet Using Plasma Actuators.” International Journal of Aeroacoustics, Vols. 5–6, 2011, pp. 635–658.
[19] Kim, J.-H., Kearney-Fischer, M., Samimy, M., and Gogineni, S. “Far-Field Noise Control in Supersonic Jets
Using Conical and Contoured Nozzles.” ASME Journal of Engineering for Gas Turbine and Power, Vol. 133,
2011, p. 081201. https://doi.org/10.1115/1.4002811.
[20] Kuo, C.-W., Cluts, J., and Samimy, M. “Exploring Physics and Control of Twin Supersonic Circular Jets.”
AIAA Journal, Vol. 55, No. 1, 2017, pp. 68–85.
[21] Yugulis, K., Hansford, S., Gregory, J., and Samimy, M. “Control of High Subsonic Cavity Flow Using Plasma
Actuators.” AIAA Journal, Vol. 7, 2013, pp. 1542–1554. https://doi.org/10.2514/1.J052668.
[22] Webb, N., and Samimy, M. “Control of Supersonic Cavity Flow Using Plasma Actuators.” AIAA Journal,
Vol. 55, 2017, pp. 3346–55. https://doi.org/10.2514/1J055720.
[23] Edgington-Mitchell, D. “Aeroacoustic Resonance and Self-Excitation in Screeching and Impinging
Supersonic Jets.” International Journal of Aeroacoustics, Vol. 18, Nos. 2–3, 2019, pp. 118–188.
https://doi.org/10.1177/1475472X19834521.
[24] Esfahani, A., Webb, N., and Samimy, M. Control of Coupling in Twin Rectangular Supersonic Jets. In AIAA
Aviation Forum, No. 2021–2122, Virtual Event, 2021, pp. 1–20.
[25] Esfahani, A., Webb, N., and Samimy, M. Coupling Modes in Supersonic Twin Rectangular Jets. In SciTech
2021 Forum, No. 2021–1292, Virtual Event, 2021, pp. 1–21.
[26] Schmidt, O., and Colonius, T. “Guide to Spectral Proper Orthogonal Decomposition.” AIAA Journal, Vol. 58,
No. 3, 2020, pp. 1023–1033. https://doi.org/10.2514/1.J058809.
[27] Samimy, M., Webb, N., and Esfahani, A. “Reinventing the Wheel: Excitation of Flow Instabilities for Active
Flow Control Using Plasma Actuators.” Journal of Physics D: Applied Physics, Vol. 52, No. 354002, 2019,
pp. 1–16. https://doi.org/10.1088/1361-6463/ab272d.
[28] Norum, T. “Screech Suppression in Supersonic Jets.” AIAA Journal, Vol. 21, No. 2, 1983, pp. 235–240.
https://doi.org/10.2514/3.8059.
[29] Tam, C., and Tanna, H. “Shock Associated Noise of Supersonic Jets from Convergent-Divergent Nozzles.”
Journal of Sound and Vibration, Vol. 81, No. 3, 1982, pp. 337–358.

12
American Institute of Aeronautics and Astronautics

You might also like