You are on page 1of 20

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

net/publication/305693111

Turbulent jet manipulation using two unsteady azimuthally separated radial


minijets

Article  in  Proceedings of The Royal Society A Mathematical Physical and Engineering Sciences · July 2016
DOI: 10.1098/rspa.2016.0417

CITATIONS READS

23 210

4 authors, including:

He Yang R.M.C. So
Harbin Institute of Technology The Hong Kong Polytechnic University
5 PUBLICATIONS   64 CITATIONS    419 PUBLICATIONS   6,380 CITATIONS   

SEE PROFILE SEE PROFILE

Yang Liu
The Hong Kong Polytechnic University
48 PUBLICATIONS   445 CITATIONS   

SEE PROFILE

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

Traditional Chinese medicine and molecular signaling View project

Lattice Boltzmann Method View project

All content following this page was uploaded by R.M.C. So on 04 October 2016.

The user has requested enhancement of the downloaded file.


Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

Turbulent jet manipulation


rspa.royalsocietypublishing.org using two unsteady
azimuthally separated
Research
radial minijets
H. Yang1,3 , Y. Zhou1,2 , R. M. C. So3 and Y. Liu3
Cite this article: Yang H, Zhou Y, So RMC,
Liu Y. 2016 Turbulent jet manipulation using 1 Institute for Turbulence-Noise-Vibration Interactions and Control,
two unsteady azimuthally separated radial Shenzhen Graduate School, Harbin Institute of Technology,
minijets. Proc. R. Soc. A 472: 20160417. Shenzhen, People’s Republic of China
http://dx.doi.org/10.1098/rspa.2016.0417 2 Digital Engineering Laboratory of Offshore Equipment, Shenzhen,

People’s Republic of China


Received: 5 June 2016 3 Department of Mechanical Engineering, The Hong Kong
Accepted: 23 June 2016
Polytechnic University, Kowloon, Hong Kong
HY, 0000-0002-1540-1793
Subject Areas:
fluid mechanics, mechanical engineering, The active manipulation of a turbulent round jet is
energy experimentally investigated based on the injection of
two radial unsteady minijets, prior to the issue of
Keywords: the main jet. The parametric study is conducted for
active jet control, unsteady minijet, the mass flow ratio Cm of the minijets to the main
flapping jet, asymmetric excitation jet, and the ratio fe /f0 of the minijet frequency to
the preferred-mode frequency of the main jet. It is
found that the decay rate of the jet centreline mean
Author for correspondence: velocity could be greatly increased if the two minijets
Y. Zhou are separated azimuthally by an angle θ = 60◦ , instead
e-mail: zhouyu@hitsz.edu.cn of by θ = 180◦ . This increase is a consequence of
the flapping motion of the jet column, and the
formation process and generation mechanism of this
flapping motion are unveiled by careful analysis of the
experimental data.

1. Introduction
The study of jet control has been of great interest over
the past few decades due to a wide range of industrial
applications and global environmental problems such
as mixing enhancement, noise suppression, chemical
reactions, heating or combustion chambers, pollutant
dispersal, spraying, and electronic equipment cooling
and drying. The rate of jet mixing plays a crucial
role in most of these applications. Clearly, this rate
is strongly influenced by the coherent structures and

2016 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

can be manipulated either passively or actively. Passive control techniques include non-
2
circular nozzles [1,2] or tabs [3], while active control strategies comprise acoustic excitation [4],
piezoelectric actuators [5,6], plasma actuators [7], synthetic jets [8] or steady/unsteady jets [9].

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
Considerable interest has been given to the fluidic means for jet mixing control (e.g. [10,11]).
The injection of the steady control jets produces a significant reduction in the potential core
and greatly increases entrainment and mixing in a low-speed jet. New & Tay [12] demonstrated
that the counter-rotating vortex pairs, generated by the penetration of continuous control jets,
contributed to jet mixing enhancement. The steady minijets also have a great effect on the
high-speed jet, producing mixing enhancement and noise suppression due to the presence
of streamwise vortices, which disrupt the generation of azimuthally large-scale coherent
structures [13].
Recently, unsteady injections are becoming popular because they can be more efficient for
the manipulation of jet mixing compared with steady injection [14]. The pulsed jets seek to
manage not only the large-scale changes via penetration but also the generation and interactions
of primary vortex rings via periodic excitation. Given a proper excitation frequency, unsteady
perturbations may maximize their effect on the entrainment characteristics of jets [15]. For a
low-speed jet, Raman [16] investigated the mixing characteristics of a rectangular jet with the
excitation of two unsteady fluidic injections and achieved a 35% reduction in the potential core
length and about a 60% increase in the normalized mass flux with respect to the unexcited jet.
Zhang [17] studied in detail the active control of a turbulent round jet using two symmetrically
arranged radial unsteady minijets, and found a much more efficient control than using steady
minijets. Three types of coherent structures were identified, i.e. the distorted vortex ring, two
pairs of azimuthally fixed streamwise vortices and sequentially ejected mushroom-like counter-
rotating structures, and jet entrainment was found to be dictated by the interactions of the three
distinct coherent structures. For compressible jets, Ibrahim et al. [18] experimentally manipulated
compressible round jets using 12 microjets equally separated along the circumference of the
nozzle exit and found a higher spreading rate, in terms of the decay of jet centreline velocity,
than steady injection.
Asymmetric disturbances are less stable than symmetric ones, due to more complex vortex
topologies [1,18], and hence may achieve better jet mixing enhancement. Longmire & Duong [19]
introduced the combination of both asymmetric passive technique and axial periodic excitation
to manipulate a round jet by attaching a stepped trailing edge and a speaker upstream of the jet
plenum. The vortex ring in the jet developed an inclined component and became asymmetrical,
resulting in significant enhancement of entrainment and spreading. Pothos & Longmire [20]
investigated a turbulent rectangular jet under the asymmetric forcing of a piezoelectric actuator.
They found that the periodic forcing at low frequencies developed asymmetric vortical structures,
causing a fast decay of the centreline velocity and a high spreading rate. Tamburello &
Amitay [21] investigated a turbulent round jet under the asymmetric excitation of a steady minijet,
pointing perpendicularly to the main jet, thus effectively accelerating the decay of the centreline
streamwise mean velocity, and enhancing spreading and increasing jet width. It is well known
that the excitation could be made unsteady to optimize the actuator efficiency (e.g. [22]). Unsteady
excitation allows jet penetration and spread to be enhanced under specific conditions [14] and
may capitalize not only on large-scale changes through penetration but also on the excitation
frequency to manipulate the inherent instabilities of the main jet. Thus, one naturally wonders
how jet mixing is enhanced under the asymmetric excitation of unsteady minijets that can provide
both asymmetric and periodic excitations.
Following the manipulation of a turbulent round jet using two opposing radial minijets by
Zhang et al. [23], we have manipulated the same jet using two asymmetrically arranged radial
minijets, separated by an angle of θ = 60◦ and 120◦ , respectively. We found that the excitation for
the case of θ = 60◦ is more efficient than that for the case of θ = 120◦ and the physical mechanisms
behind them are essentially the same. Therefore, this work is focused on the case of θ = 60◦ . One
objective is to understand the dependence of the jet decay rate on control parameters, i.e. the
mass flow ratio (Cm ) of the minijets to the main jet and the frequency ratio fe /f0 , where fe is the
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

y, V
3
L = 17 q = 60°
ds = 0.9

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
y, V minijet
Vj
r Vr

x, U j
z, W
main jet Ue
D = 20

dr = 1 stationary disc

rotating disc

Figure 1. Schematic of minijet assembly.

excitation frequency of minijets and f0 is the preferred-mode frequency of the main jet. Another
objective is to gain insight into the physical mechanism behind possible effective control of the
main jet under asymmetric manipulation. The experimental set-up is described in §2. Section 3
presents the dependence of jet decay rate on Cm and fe /f0 , and a comparison of the performance
of the asymmetric and symmetric controls. The minijet actuator performance and its effect on the
initial condition of the main jet are discussed in §4. The effect of Cm and fe /f0 on the flow structure
is investigated in §5, while the evolution of the flow structure is examined in §6, along with
a discussion of the physical mechanism behind the effective jet manipulation. The conclusions
drawn are presented in §7.

2. Experimental details
Experiments on an axisymmetric air jet were performed. The jet facility consists of a main jet
with different minijet assemblies, as described by Zhang et al. [23]. The flow rates of the main
jet and minijets are adjusted via two separate control valves and measured using two separate
flow meters. The air supplied for the main jet passes through a mixing vessel, mixed with seeding
particles in the case of flow visualization or particle image velocimetry (PIV) measurements, and a
plenum box before entering the main jet chamber. The chamber includes a 300 mm long diffusing
section, a cylindrical part of 400 mm in length and a smooth contraction nozzle. The inlet diameter
of the contraction section is 114 mm and the exit diameter D is 20 mm, while the profile of the
nozzle contraction is given by R = 57 − 47 sin1.5 (90 − 9x /8) following closely the design of Mi
et al.’s nozzle [24]. The contraction was extended with a 47 mm long smooth tube of diameter D.
The jet Reynolds number ReD based on D is fixed at 8000. The main jet is issued with a ‘top-hat’
mean velocity distribution and a low turbulence intensity of 0.25%. The minijet assembly includes
a stationary and a rotating disc (figure 1). The stationary disc is made with two orifices of 0.9 mm
in diameter, separated azimuthally by 60◦ degrees, located at 17 mm upstream of the main jet exit.
The rotating disc has 12 orifices of 1 mm in diameter and is driven by a servomotor; the orifices
are equally separated azimuthally. Once the orifices on both discs are aligned during rotation,
two pulsed minijets are injected into the main jet. Two control parameters are examined, i.e. the
ratio fe /f0 of the minijet excitation frequency fe to the main jet preferred-mode frequency f0 , and
the mass flow ratio Cm of the minijets to the main jet, over a range of fe /f0 = 0.31–1.24 with Cm
varying from 1.1% to 10.9%.
The origin of the coordinate system is chosen at the centre of the main jet exit, with the x-axis
positioned along the streamwise direction. The corresponding cylindrical coordinate system
(x, ϕ, r) is defined in figure 1. Measurements were conducted in the symmetry plane, i.e. the (x, y)
plane, of the two minijets, the orthogonal plane through the jet geometric centreline, i.e. the (x, z)
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

plane, and a number of cross-sectional (y, z) planes using flow visualization, PIV and hotwire
4
techniques. Following Zhou et al. [11], the decay rate K of the jet centreline mean velocity is used
to evaluate jet mixing, given by K = (Ue − U5D )/Ue , where Ue and U5D denote the jet centreline

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
time-averaged velocities at x∗ = x/D = 0 and 5, respectively.
The streamwise velocity was measured using a single calibrated hotwire probe, operated on a
constant temperature circuit (Dantec Streamline) with an overheat ratio of 1.8. The tungsten wire
was 5 µm in diameter and approximately 1.5 mm in length. The bridge output voltage signal was
offset, amplified and low-pass filtered at a cut-off frequency of 3 kHz before being digitized at
a sampling frequency of 6 kHz with a 12-bit A/D converter. The duration for each record was
80 s. The hotwire signal was calibrated based on the streamwise mean velocity measured with
a standard static Pitot tube, connected to an electronic micromanometer (Furness FCO510) and
placed side by side with the hotwire near the centre of the nozzle exit.
Flow visualization and PIV measurements were carried out using a Dantec standard PIV 2100
system. A seeding particle generator (TSI 9307-6) was used to generate smoke, which was mixed
with air in the mixing vessel. Smoke-seeded flow was illuminated by a laser sheet generated
from two New Wave standard pulsed laser sources of 532 nm in wavelength with a maximum
energy output of 120 mJ per pulse. Each laser pulse lasted for 0.01 µs. Instantaneous images were
obtained at a sampling rate of 4 Hz using a CCD camera (HiSense MkII, double frames, 2048 ×
2048 pixels). A Dantec Flow Map Processor (PIV 2001 type) was used to synchronize illumination
and image capturing. The settings of the PIV system are chosen to be the same for both flow
visualization and PIV measurements.

3. Dependence of jet decay rate on Cm and fe /f0


The control performance depends on the mass flow ratio, excitation frequency ratio, number,
shapes and geometric arrangement of unsteady minijets [6,9]. In this work, the effect of Cm and
fe /f0 on jet entrainment was investigated. The main jet velocity is 6 m s−1 and the preferred-mode
frequency is 128 Hz, corresponding to a Strouhal number of StD = f0 D/Ue = 0.43, which falls in
the range (0.24–0.64) reported by Gutmark & Ho [25]. Figure 2 presents the dependence of the
jet centreline mean velocity decay rate (K) on Cm and fe /f0 , along with that under symmetric
control when θ is 180◦ and the injections of two minijets are in-phased, which serves as a
reference for comparison. The K variation for these two cases is qualitatively similar. However,
the asymmetric control case is significantly more effective in enhancing jet entrainment than the
symmetric case; for example, K is increased by 30% at fe /f0 = 1.0 for Cm = 1.6% when compared
with the symmetric case (figure 2a). A local K maximum occurs at fe /f0 = 1.0, as noted by others
(e.g. [7]). The dependence of K on Cm at fe /f0 = 1.0 (figure 2b) is characterized by a rapid rise in
K with initial increase in Cm , reaching a maximum K (0.284) at Cm = 2.0% and then dropping to
0.186 at Cm = 4.2%. A further increase in Cm leads to K rising asymptotically to a constant. This
constant is apparently substantially larger for the asymmetric case compared with the symmetric
case. The case of Cm = 2.0% and fe /f0 = 1.0 only gives rise to a local maximum K. From the
practical application point of view, a manipulation should ideally require the minimum energy
consumption [9,26]; in other words, the mass flow of the minijets is required to be as small
as possible. With this consideration, a minijet excitation with Cm = 2.0% and fe /f0 = 1 gives the
optimal condition.
Note that the injection time of the minijets (or disc rotation frequency) is one of the factors
that affect the character of the jet in crossflow [27]. For a fixed minijet mass flow rate, the rotation
frequency may have an effect on the minijet velocity. As fe /f0 → 0, the injected flow may exhibit
a behaviour similar to that of the steady injection. On the other hand, as fe /f0 → ∞, the injection
time of each pulse approaches 0 and accordingly the injected mass goes to zero, that is its velocity
approaches zero. As such, the minijet velocity depends on how fast the openings in the rotor pass
those in the stationary disc, i.e. the opening time of the orifice issuing the minijet. The main jet
exit velocity is 6 m s−1 if unexcited; as Cm increases, the main jet exit velocity changes little in the
range of Cm tested (not shown). Two factors may contribute to this observation. First, the actual
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a) (b)
0.4 0.4 5
q = 60°, Cm = 1.6%
q = 180°, Cm = 1.6%

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
0.3 natural jet (Cm = 0, fe/f0 = 0) 0.3

K 0.2 0.2

q = 60°, fe/f0 = 1
0.1 0.1 q = 180°, fe/f0 = 1
natural jet (Cm = 0, fe/f0 = 0)
0 0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0 2 4 6 8 10 12 14
fe/f0 Cm (%)

Figure 2. Dependence of K on (a) fe /f0 (Cm = 1.6%) and (b) Cm (fe /f0 = 1), when compared with the case of θ = 180◦ .

mass flow rate of the minijets injected into the main jet is less than the measured mass flow rate
due to a clearance, necessary to allow the rotor to rotate, between the stationary and rotating discs,
which implies a flow leakage. This leakage is worse at a large Cm . Second, there is an appreciable
increase in the jet width measured at x∗ = 0.05 (not shown) for large Cm .

4. Performance and role of the minijet actuator


The minijet actuators can increase the jet decay rate substantially (figure 2). Therefore, it is
very important to carefully examine the flow physics in order to explore the minijet actuator
performance and its effect, especially at the optimal excitation condition (Cm = 2.0%, fe /f0 = 1.0).
In order to study actuator performance, an experiment was carried out by operating one minijet at
a mass flow rate and a frequency corresponding to Cm = 0.99% and fe /f0 = 1.0 in the absence of the
main jet. A single calibrated hot wire was deployed at x∗ = −0.85 (ϕ = 120◦ and r∗ = r/D = 0.35),
where the minijet was injected. The wire is oriented normally to the axis of the minijet orifice. The
time history of the radial fluctuating velocity −vr signal (figure 3) shows periodically uniform
peaks, thus suggesting the actuator can produce the observed uniform minijet flow. Furthermore,
to explore the effect of the minijets on the initial condition of the main jet, two minijets are
operated at Cm = 2.0% and fe /f0 = 1.0 in the presence of the main jet. Following Zhou et al. [11], a
hotwire probe is placed at x∗ = −0.7 (ϕ = 120◦ and r∗ = 0.35), almost immediately downstream
of one of the minijets. Compared with the natural jet (Cm = 0 and fe /f0 = 0), the normalized
streamwise velocity u/Ue signal (figure 4a) exhibits significant fluctuations in the excited jet, thus
indicating the minijet actuators can produce large amplitude perturbations. This is consistent
with the result obtained by Choi et al. [28] in their investigation of a supersonic jet excited with
16 unsteady microjets. The presence of two concave peaks apparently reflects the interaction
between two approaching minijets. At θ = 120◦ –180◦ , only one concave peak is observed for each
pulse, indicating that the two minijets do not interact appreciably with each other. As θ decreases
to 60◦ , the two minijets are more close to each other and more mutually aligned, hence potentially
interacting with each other. As a result, the flow structure of the first minijet (located just upstream
of the hotwire) may be affected by the presence of the second, thus leading to the appearance of
two concave peaks in the hotwire signal. At the extreme case θ = 0◦ , the two minijets merge into
one, forming a single high-momentum jet. The corresponding spectrum Eu is shown in figure 4b.
There is no obvious peak in the natural jet except that at f ∗ = fD/Ue = 0.17 (f = 50 Hz), which
is associated with the noise linked to the power supply. Once excited, Eu exhibits a number
of pronounced peaks at f ∗ = 0.43 (fe = 128 Hz) and higher harmonics due to periodic excitation
(figure 4b), as previously reported by Raman & Cornelius [26]. Evidently, the unsteady minijet
injections can produce velocity fluctuations with large enough amplitude around the required
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

16
t = 1/feª 7.81 ms 6

12

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
–vr (m s–1)
8

0 0.005 0.010 0.015 0.020 0.025 0.030


T (s)

Figure 3. Time history of −vr produced by one pulsed minijet, measured at x ∗ = −0.85 (r∗ = 0.35 and ϕ = 120◦ ).
One minijet operated at Cm = 0.99% for fe /f0 = 1.0 in absence of main jet.

(a) 1.2 (b)


10–1 StD = 0.43 2StD (2fe)
1.0 ( fe = 128 Hz) 3StD (3fe)
10–2
0.8
10–3
u/Ue

0.6
Eu

10–4
0.4 10–5
natural jet natural jet
0.2 10–6
manipulated jet (Cm = 2.0%, fe/f0 = 1) manipulated jet (Cm = 2.0%, fe/f0 = 1)
0 10–7 –2
0 0.005 0.010 0.015 0.020 0.025 0.030 10 10–1 1 10
T (s) f* = fD/Ue

Figure 4. (a) Time histories and (b) power spectral density function Eu of u obtained at x ∗ = −0.7 (ϕ = 120◦ and r∗ = 0.35),
downstream of one of the minijets.

frequency at the upstream of the main jet exit and the subsequent shear layer issuing from the
nozzle. Such periodic velocity fluctuations may accelerate the growth of the shear layer, forming
large-scale coherent structures [29]. Therefore, the unsteady minijet excitation may cause the shear
layer at the actuator side to roll up into vortices earlier than that in the opposite side of the
actuators.
To understand the effect of Cm on the initial state of the main jet exit, u∗rms = urms /Ue was
measured at x∗ = 0.05 in the natural jet case as well as in the manipulated jet case for various Cm
(figure 5). In the natural jet case, the small humps of u∗rms occur at y∗ = y/D ≈ ±0.46, resulting
from the instability wave in the initial shear layer. Once the natural jet is manipulated, u∗rms
changes substantially. At small Cm (=2.0%), u∗rms does not show any appreciable increase for
y∗ < 0 but does for y∗ ≥ 0, apparently as a result of minijet injections. The major characteristics of
the jet remain unchanged except for the flow near the actuator side. The hump at y∗ ≈ 0.46 grows,
suggesting the fluctuation in the shear layer is amplified, which may account for the rapid growth
of the shear layer. Considerable increase in u∗rms along the y-axis with a maximum occurring
around the centre is observed once Cm is increased to 4.2%. It could be inferred that the two
minijets may have penetrated the jet centre, clashing with each other at the jet exit. The turbulence
produced by the collision of two minijets may also account for the increase in u∗rms along the z-axis,
especially around the centre (figure 5b). As Cm is increased to 9.3%, u∗rms increases throughout y∗
and z∗ (= z/D), especially for y∗ < 0 (figure 5a) and even z∗ < −0.2 (figure 5b), suggesting that the
two minijets may collide with each other even before reaching the nozzle exit. The resultant mixed
fluid travels along the negative y-axis, producing a highly turbulent jet right from the beginning.
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a) (b) 7
0.4 0.4
natural jet
Cm = 2.0%, fe /f0= 1.0
Cm = 2.0%, fe /f0= 1.0

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


Cm = 4.2%, fe /f0= 1.0

...................................................
0.3 Cm = 4.2%, fe /f0= 1.0 0.3
urms/Ue Cm = 9.3%, fe /f0= 1.0
Cm = 9.3%, fe /f0= 1.0
0.2 0.2

0.1 0.1

0 0
–0.6 –0.4 –0.2 0 0.2 0.4 0.6 –0.6 –0.4 –0.2 0
y* z*

Figure 5. Radial distribution of turbulent intensity urms /Ue at x ∗ = 0.05 of the manipulated jet in (a) x–y plane and
(b) x–z plane.

Another consideration of the minijet actuator is the effect of the rotating disc; it might produce
azimuthal velocity to the main jet. Firstly, the rotating disc may have an effect on f0 . Consider
the optimal excitation condition ( fe /f0 = 1.0, Cm = 2.0%). Attention is focused on the jet with and
without the inner disc rotating at the speed corresponding to fe /f0 = 1.0 and Cm = 0. According
to Hussain [30], f0 is defined by the predominant vortex passage frequency near the end of the
potential core. It is found that f0 decreases from 143 Hz (StD = 0.48) without the disc rotating to
128 Hz (StD = 0.43) with the disc rotating (not shown). The reason for the frequency decrease
could be attributed to the fact that the rotating jet may result in a slight growth of the axial shear
layer thickness [31], and the preferred-mode frequency is inversely proportional to the initial
shear layer momentum thickness (e.g. [32]). For clarity, f0 is equal to 128 Hz in the natural jet set-
up, unless otherwise stated. Furthermore, the K value increases from 2.4% to 5.4%. Finally, it is
necessary to examine whether centrifugal instability due to swirling effect occurs in the present
natural jet/minijet set-up. The PIV-measured azimuthal mean velocity at x∗ = 0.1 is very small,
only discernible at r∗ ≈ 0.5 with a maximum value of about 0.06Ue (not shown). This suggests,
following the criterion of Oberleithner et al. [31], that the rotation of the inner disc is not strong
enough to excite the centrifugal instability that occurs in a swirling jet. In summary, the major
characteristics of the natural jet remain largely unchanged with the disc rotating.

5. Effect of Cm and fe /f0 on flow structure


From the current experiments on perturbed jets, it can be seen that Cm and fe /f0 have a great effect
on the jet decay rate (figure 2). This could be explained from the jet flow structures under the
effect of different Cm and fe /f0 . Thus, in this section, the flow structures under typical excitation
conditions are examined through the velocity signals and photographs of flow visualization.
The evolutions of the hotwire-measured streamwise velocity u signals are shown in figure 6,
with and without manipulation. In the natural jet (Cm = 0 and fe /f0 = 0), the signal shows a quasi-
periodically sinusoidal wave at x∗ = 2.0, indicating the shear layer rolls up, forming vortices due
to Kelvin–Helmholtz instability [32]. Then, the frequency of large events appears to be halved at
x∗ = 3.0, probably resulting from vortex pairing [29,33]. Once manipulated, the signals exhibit a
dramatic change in characteristics, depending on Cm . A number of observations can be made from
the u-signals at three representative Cm in figure 6. Firstly, at Cm = 2.0%, u signals exhibit early
sign of periodic fluctuations at given y∗ and z∗ , suggesting an early roll-up of the shear layer.
The signals remain smooth or laminar even at x∗ = 2.0, especially along y∗ = −0.3 or z∗ = 0.3.
Secondly, given Cm = 9.3%, all the signals, be they from the x–y or x–z plane, appear fluctuating
randomly even at x∗ = 0.5, suggesting a turbulent state from the very beginning. This observation
is further supported by flow visualization data (figure 7a). Note that larger spreading can be
observed in the x–y plane than in the x–z plane, resulting from the more effective entrainment by
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

natural jet manipulated jet manipulated jet manipulated jet 8


(Cm = 0, fe /f0= 0) (2.0%, 1) (4.2%, 1) (9.3%, 1)
x* (a) (b) (c) (d)

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
0.5

1.0
y* = 0.3

1.5

2.0

3.0

0.5

1.0
y* = –0.3

1.5

2.0

3.0

0.5

1.0
z* = 0.3

1.5

2.0

3.0

Figure 6. Typical signals of streamwise velocity u. Duration is 0.05 s.

turbulent vortical structures [11]. Thus, this turbulence, boosted by increasing Cm , is responsible
for the increasing K with Cm > 4.2% (figure 2b). This mechanism is apparently different from that
observed when the jet is excited by a Cm in the range 0 < Cm < 3.5%, which will be discussed
in detail later. Finally, at Cm = 4.2%, the signals obtained at y∗ = 0.3 and −0.3 both exhibit the
transitional characteristics, though that at z∗ = 0.3 remains smooth or laminar until x∗ = 1.5. This
is confirmed by flow visualization data (figure 7b). The laminar vortices can only be seen at
x∗ ≈ 1.2 in the x–z plane, while a turbulent state occurs in the x–y plane. The observations are
consistent with the distribution of urms at the jet exit (figure 5).
The observations at Cm = 2.0% are particularly interesting because of the impressive
enhancement of the jet entrainment rate with such a small Cm . One naturally wonders what
physical mechanism is responsible for this behaviour. Understanding of this physical behaviour
can be gleaned from the flow visualization data presented in figure 7c,d. The shear layer in the
natural jet starts to roll up at x∗ ≈ 2 and the vortices formed display symmetry about the centreline
(figure 7d) due to the occurrence of the well-known ring vortices. Under the manipulation of
Cm = 2.0%, there is a great change in the flow structure (figure 7c(i)(ii)). Firstly, the occurrence
of shear layer roll-up moves upstream at x∗ ≈ 1. Secondly, the large-scale vortices appear very
different between the x–y and the x–z plane. While the vortices in the x–z plane remain almost
symmetrical about the centreline, those in the x–y plane do not. This observation is derived
from the visualization result showing that two minijets excitation is symmetrical about the x–y
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a) (b) (c) (d)


9
(i) (i) (i)
5

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
4
x–y plane

3 x*

(ii) (ii) (ii)


5

4
x–z plane

3 x*

Figure 7. Photographs of typical flow structures from flow visualization. (a) Manipulated jet (Cm = 9.3%, fe /f0 = 1),
(b) Cm = 4.2%, fe /f0 = 1, (c) Cm = 2.0%, fe /f0 = 1 and (d) natural jet (Cm = 0, fe /f0 = 0).

plane, but not the x–z plane. This is in agreement with previous investigations on asymmetric
manipulation of the jet, where asymmetric excitation may lead to asymmetrically arranged
vortices [19,20]. Finally, the jet column in the x–y plane wobbles rightwards at x∗ ≈ 2 and then
leftwards at x∗ ≈ 2.7 and rightwards again further downstream, along with a significant spread
compared with that at x∗ ≈ 2. Such an oscillation of the jet column is similar to observation
in a round jet excited by two anti-phased slot jets at a Reynolds number of 3600 (see Figs 6
and 7 in [22]), which was referred to as the flapping jet column. As a matter of fact, the phase
spectrum Φu1u2 between u1 and u2 measured at y∗ = −0.3 and 0.3 (x∗ = 2), respectively, shows
a jump from 0 in the natural jet to near π or −π over a broad range of frequencies about the
preferred-mode frequency StD = 0.43 (figure 8). Note that both Φu1u2 = −π and π correspond
to anti-phase. It may be inferred that, unlike the natural jet, the flow is oscillating while moving
downstream. Goldschmidt & Bradshaw [34] proposed the flapping of the jet based on the negative
correlation between two fluctuating streamwise velocities obtained on opposite sides of the jet.
Figure 9a presents two snapshots of typical instantaneous lateral velocity vectors captured at
x∗ = 2.0 using PIV, which are selected to present the highly repeatable velocity vectors, showing
the predominantly upward and downward motions, respectively, and the unequivocal evidence
for the flapping motion of the jet column (figure 7c(i)). Note that the jet spreads along the flapping
direction. Thus, it may be inferred that the flapping motion is responsible for the increased
spreading and the rapid decay of the jet. Considering the presence of an initially tilting vortex
ring and the absence of an axisymmetric counterpart, jet flapping may be due to the excitation
of the m = ±1 mode, where m denotes the azimuthal mode [6,7]. The momentum thickness of
the natural jet is θm = 0.17 mm, thus giving θm /D = 0.0085. Previous investigations [35–37] have
confirmed that the initial region of an axisymmetric jet surrounded by a thin initial shear layer
(θm /D 1) is equally unstable to an axisymmetric (m = 0) as well as to the first helical modes
(m = + or −1). Samimy et al. [7] achieved a higher enhancement of jet mixing and a faster decay
rate of jet centreline velocity by exciting a round jet with the m = ±1 mode at the jet preferred-
mode frequency than that excited with the m = 0 mode at the same frequency, where the m = 0, ±1
modes are forced by changing the way actuators are operated. They observed that the structures
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

natural jet 10
4
manipulated jet (Cm = 2.0%, fe /f0 = 1)

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
2

F u1u2
0

–2

–4
StD = 0.43

0 0.5 1.0 1.5 2.0


f/f0

Figure 8. Phase spectrum Φu1 u2 between two streamwise velocity signals measured at x ∗ = 2 and y∗ = ±0.3.

(a) (i) (ii)

–1.0 3 m s–1

–0.5

y*
0

0.5

1.0

–1.0 –0.5 0 0.5 1.0 –1.0 –0.5 0 0.5 1.0


z* z*
(b) a(ii) a(i)
y flow direction

time

lateral velocities

Figure 9. (a) Two typical PIV snapshots of instantaneous velocity vectors in a cross-sectional plane of the manipulated jet
(Cm = 2.0%, fe /f0 = 1) at x ∗ = 2.0. (b) Schematic of the jet flapping motion. The a(i) and a(ii) in panel (b), corresponding to
(a(i)) and (a(ii)), respectively, represent two typical phases in the process of the jet flapping motion.

are symmetric in the m = 0 mode excitation, but staggered in the m = ±1 mode excitation. Parekh
et al. [6] employed two anti-phase slot jets to excite the m = ±1 mode of a round jet, thus
producing a flapping jet that enhanced jet mixing effectively. Although the excitation techniques
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a) (b) (c) 11


5

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
4

3 x*

Figure 10. Photographs of typical flow structures from flow visualization in the x–y plane of the manipulated jet (Cm = 1.6%):
(a) fe /f0 = 0.85, (b) 1.0, (c) 1.24.

are different, the present excitation (Cm = 2.0%, fe /f0 = 1) also introduces an asymmetric excitation
at the preferred-mode frequency, which probably excites the unstable m = ±1 mode, producing
the staggered vortices and eventually a flapping jet.
The K value experiences an obvious sharp rise as fe /f0 approaches 1 (figure 2a). On the other
hand, the drop in K is not so sharp beyond fe /f0 = 1. The distinct behaviours suggest a difference
in the flow physics involved. Figure 10 presents the typical flow visualization photographs taken
in the x–y plane of the excited jet at fe /f0 = 0.85, 1.0 and 1.24 for a fixed Cm = 1.6%, respectively.
The case of fe /f0 = 1.0 serves as a reference. The excited jet reveals staggered vortices and a
vigorous flapping motion. By contrast, if forced at fe /f0 = 0.85, the vortices appear approximately
symmetric about the centreline, and the flapping motion is not readily visible (figure 10a). The
absence of jet flapping may lead to weak spreading, and hence a relatively small K. At fe /f0 = 1.24,
the flow structure is quite similar to that at fe /f0 = 1.0. However, with increasing excitation
frequency, the time interval between two adjacent vortices is reduced; thus, relatively small
vortices emerge with reduced spatial separation [20,33,38], which may weaken the induced lateral
velocity due to vortex pairing. As a result, jet flapping is less pronounced and the flow spreads
less (figure 10c) than that at fe /f0 = 1.0 (figure 10b), leading to a drop in K. Iio et al. [39] also
confirmed that excitation at fe /f0 > 1 can produce a weak flapping motion of the jet column. We
may conclude that fe /f0 ≥ 1 could create the right condition to generate flapping motion in a round
jet, though this motion is most pronounced at fe /f0 = 1.

6. Flow structure development


A flapping jet is generated under the excitation of Cm = 2.0% and fe /f0 = 1, which may account
largely for the increased jet decay rate and is of great interest. To further understand the
flow structure development under this excitation, flow visualization and PIV measurements
were conducted in the cross-sectional planes and the streamwise planes, with and without
manipulation (figures 11–15).
The initial formation of vortex ring in the manipulated jet may be investigated from flow
photographs, taken in cross-sectional planes from x∗ = 0.7–1.3, along with their counterparts
without manipulation (figure 11). The flow images of the natural jet look like a full moon
(figure 11a(i)(iii)), which is the potential core of the jet. Apparently, there is no roll-up of shear
layers. Once perturbed, the shear layer on the upper side always rolls up first at x∗ = 0.7 to form a
vortical structure (figure 11b(i)). This is due to the unsteady minijet injections placed on the upper
side. The periodic excitation may accelerate shear layer growth [19]. At x∗ = 1.0, the shear layer
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

x* = 0.7 x* = 1.0
12
part of the ring structure

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
(a(i)) (b(i)) (a(ii)) (b(ii))

x* = 1.3

vortex ring streamwise structures vortex ring

(a(iii)) (b(iii)) (b(iv)) (b(v))


z

Figure 11. Typical photographs of flow visualization in the cross-sectional planes. (a) Natural jet, (b) manipulated jet
(Cm = 2.0%, fe /f0 = 1). Flow is outward from the paper.

on the lower side also starts to roll up (figure 11b(ii)). The black curve between the vortex and
the potential core is the ambient fluid entrained during vortex formation. Both upper and lower
vortical structures are part of slightly tilting vortex rings, which will be elaborated later. At x∗ =
1.3, complete ring vortices are captured, though with small distortions (figure 11b(iii)(v)). Also,
the braid region between ring vortices is shown in figure 11b(iv), two streamwise vortices are
discernible in this image, most probably generated under the high strain rate of the braid region.
In order to explore the downstream evolution of vortex rings, four images in the x–y plane are
selected to present typical phases of the vortex evolution (figure 12a). We wish to point out that
the flow behaviour observed is highly repeatable. For ease of discussion on the two oppositely
orientated ring vortices, the vortices are designated as ring vortex A and ring vortex B as shown
in figure 12a. The structure B0+ moves inwards to B1+ as a consequence of induction of the
downstream vortex. Consequently, vortex ring B0 rotates clockwise to B1 around an axis parallel
to the z-axis. On the other side, vortex ring A0 grows to a larger size A1. As such, the adjacent
vortex rings A1 and B1 are oppositely tilted (figure 12a(ii)(iv)), as indicated by the upper and
lower ends (A1+ and A1−) and (B1+ and B1−), respectively (figure 12a(ii)(iv)). Such a process is
sketched in figure 12b(i)(ii). Recall the vortex rings captured in the cross-sectional plane of x∗ = 1.3
shown in figure 11b(iii)(v); these could be interpreted as corresponding to B1 and A1, respectively,
and the incomplete rings, i.e. the upper turbulent segments, should be the consequence of tilting.
Further downstream, the vortex rings are more stretched and tilted. The ring vortex B rotates
clockwise from B1 to B6, while A turns anti-clockwise from A1 to A6 (figure 12a). As a result, two
adjacent rings approach each other at one end (e.g. B2+ and A4+ on the upper side, A4− and
B6− on the lower side as shown in figure 12a(i)), forming zigzag-arranged vortex structures. This
evolution of vortex tilting is sketched in figure 12b and represented by sketches in figure 12b(ii)–
b(iii). Note that two interacting vortices on either the upper or the lower side undergo localized
pairing. Specifically, on the upper side, the vortex B1+ evolves downstream to B2+, approaching
to and then pairing with A4+, thus forming AB+. On the lower side, the vortex A1− moves
upwards to A3−, catching up with B5−. The two vortices undergo pairing, forming AB−. It
should be noted that the pairing process is not symmetrical about the jet centreline, a behaviour
that is different from that observed in a natural jet. This asymmetric arrangement of the pairing
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a)
13
(i) (ii)
A4+ A1+ A5+
A0+ B2+

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
A0 A4 A1 A5
B2
B6 B3

A1–
AB– B6– A4–

(iii) AB+ (iv)


B0+ B1+

B0 A2 A6 A3
B1
B4 B5

y A2– B1– A3–


B4–
B5–

1 2 3 4 5 1 2 3 4 5
x* x*
(b)
(i) (ii) (iii)

Figure 12. (a) Photographs of typical flow structures from flow visualization in the x–y plane of the manipulated jet
(Cm = 2.0%, fe /f0 = 1). (b) Sketch to illustrate the evolution of vortex tilting. The sequence is from (b(i)) to (b(iii)).

(a) (b)
0.3 –0.25 0.2 –0.2 0.25 –0.2
1.0

y* 0

–1.0
–0.25 0.25 –0.25 0.3 –0.35 –0.4
1 2 3 4 1 2 3 4
x* x*

Figure 13. Typical v ∗ -contours in the x–y plane: (a) natural jet, (b) manipulated jet (Cm = 2.0%, fe /f0 = 1). Contour
interval = 0.05. Continuous and dashed lines represent the positive v ∗ (upward motion) and the negative (downward motion),
respectively. Symbols ‘+’ and ‘×’ denote anti-clockwise and clockwise vortices, respectively.
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

(a) x* = 2.0 x* = 2.4 x* = 3.0 x* = 5.0 14


(i) (ii) (iii) (iv)

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
(b)
(i) (ii) (iii) (iv)

mushroom-like mushroom-like
structures structures

Figure 14. Typical photographs of flow visualization in the cross-sectional planes. (a) Natural jet, (b) manipulated jet
(Cm = 2.0%, fe /f0 = 1). Flow is outward from the paper.

(a) (b)

1 (i) U* 0.1 (i) V* –0.02


natural jet

0.9 0.02
y* 0

–1 0.1

1 (ii) U* 0.1 (ii) V* 0.08


manipulated jet (Cm = 2.0%, fe /f0 = 1)

–0.02
0.9
y* 0 –0.04

–1 0.04 –0.12
0.1

1 (iii) U* (iii) W* 0.04 –0.06


0.1

0.9
z* 0 –0.02
–0.04
–1 0.1
0.06
1 2 3 4 5 6 7 1 2 3 4 5 6 7
x* x*

Figure 15. The contours of time-averaged (a) streamwise velocity and (b) lateral velocity, the contour increments being 0.1
and 0.02, respectively.
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

process is also observed in the numerical study of a bifurcating jet [40]. In that study, it was
15
reported that the jet splits into two separate jets. As the asymmetrically arranged vortex pairing
only occurs in the x–y plane rather than in the x–z plane, as is evident in figure 7c(i)(ii), this event

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
could be directly responsible for the flapping observed in the x–y plane.
To further explore the connection between the flapping motion and the asymmetrically
arranged vortex pairing, the typical instantaneous contours of the lateral velocity v ∗ = v/Ue in
the x–y plane are shown in figure 13. Symbols ‘+’ and ‘×’ denote vortex centres above and below
the centreline, respectively. Solid and broken contours represent the positive v ∗ or upward motion
and the negative or downward motion, respectively. In the natural jet, the velocity contours are
nearly symmetric about the jet centreline, ambient fluid is induced either outward or inward and
the potential core fluid does not move laterally until x∗ ≈ 4.5. Once manipulated, the jet exhibits
alternate upward and downward motions from x∗ ≈ 1.4 onwards. Ambient fluid is entrained
upstream of two interacting vortices, evident in the vorticity contours (not shown), while the core
fluid is tossed out downstream of the two vortices. Asymmetrically arranged, vortex interactions
entrain ambient fluid on the lower side and push the core fluid out on the upper side at the
same streamwise position (e.g. x∗ ≈ 1.6 in figure 13b), thus producing a strong upward motion.
Similarly, a strong downward motion occurs at x∗ ≈ 2.2 in figure 13b. As a result, a flapping
motion of the jet column is generated.
The mechanism for generating the flapping motion proposed here is distinct from those
previously reported by Reynolds et al. [41]. No pairing was observed in their investigation. In their
review of a round bifurcating jet, obtained by combining the axial and circumferential excitations
with a ratio of axial to transverse excitation frequency equal to two, Reynolds et al. [41] proposed
that the opposite tilting of adjacent vortex rings tends to tilt toward each other due to their mutual
induction. Consequently, vortex rings move away from the jet centreline and stretch the jet core
fluid between them, producing a flapping jet (refer to fig. 2 in [41] for details).
To gain better understanding of the remarkable spread (e.g. figure 12), we explore the flow
structure based on the photographs captured in a series of cross-sectional planes at x∗ = 2.0, 2.4,
3.0 and 5.0 (figure 14). The natural jet, also shown for reference, is approximately axisymmetric. At
x∗ = 2.0, the shear layer rolls up to form an axisymmetric vortex ring (figure 14a(i)). This ring starts
to become unstable at x∗ = 2.4 (figure 14a(ii)), though the jet remains essentially axisymmetric. At
x∗ = 3.0, small mushroom-like structures emerge around the vortex ring (figure 14a(iii)). These are
generated in the braid region between two consecutive vortex rings due to azimuthal instabilities
and stretching around the following vortex ring [42]. At x∗ = 5.0, the flow becomes fully turbulent
and the jet is still roughly axisymmetric (figure 14a(iv)).
Once excited, transition to fully turbulent flow occurs early, at x∗ ≈ 2.0, in the manipulated jet,
along with a significant spread along the y direction. It should be noted that, comparing with
a natural jet, early formation and subsequent pairing of the vortices occur in a manipulated
jet. During vortex pairing, a large strain rate is generated [32], causing the braid region to
deform [43] and thus producing counter-rotating streamwise vortex pairs, which stretch upstream
and superimpose on the following primary vortex [42,44]. Hence, at x∗ = 2.0–2.4, the mushroom-
like structures can be observed at the edge of the jet (figure 14b(i)(ii)). Note that such streamwise
vortices and the azimuthal vortices appear to be pushed out by the flapping motion. Owing to the
presence of jet flapping, the lateral velocity is dramatically increased in the flapping plane [45],
and thus the jet fluid is transported outwards approximately vertically. Further downstream at
x∗ = 3.0 (figure 14b(iii)), the vortices break down, and the manipulated jet looks less ordered and
becomes turbulent. The jet fluid is pushed outwards approximately along the y-axis. Terashima
et al. [46] proposed that flapping motion could increase turbulent diffusion to the outer region
of the jet. Finally, a state of fully turbulent flow occurs at x∗ = 5.0, along with a significant
spread along the y-axis (figure 14b(iv)). Note that the manipulated jet spreads only along the
flapping direction (y−axis). It may be inferred that it is the jet flapping motion that contributes
to the significantly increased spreading and hence rapid jet decay. This finding is consistent with
previous reports (e.g. [6,7,47]), where the flapping motion was found to be effective in increasing
jet spreading and large-scale mixing. The flapping jet in this study is realized with two in-phase
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

minijets separated by an angle of 60◦ , distinctly different from two anti-phase actuators deployed
16
at the opposite sides used by previous researchers, e.g. piezoelectric actuators [6], slot jets [6],
electromagnetic flap actuators [48], acoustic actuators [39,49] and synthetic jets [47].

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
The time-averaged flow field obtained from PIV measurements provide a quantitative
measurement of jet entrainment and spreading. The streamwise and lateral mean velocities
are derived from 1500 pairs of PIV images. Figure 15a presents the contours of time-averaged
streamwise velocity U∗ = U/Ue of the jet, with and without manipulation. Two observations can
be made. Firstly, under the excitation of minijets, the location, where the contour of U∗ = 0.9
intersects the jet axis, has moved from x∗ ≈ 6.1 to 3.4, implying that the excitation gives rise to
a reduced potential core length by almost one half. Secondly, the edge of the jet splits into a
larger angle than the unexcited jet in the x–y plane, as indicated by a comparison between the
two contours of U∗ = 0.1 for the manipulated and natural jets. Note that the outer boundary
of the jet (e.g. the U∗ = 0.1 contour) is pushed outward only for x∗ < 3 in the x–z plane, which
may be due to early roll-up in this plane (figure 7c(ii)). The widening of the jet is much more
pronounced in the x–y plane than in the x–z plane, indicating greater spreading in the x–y plane.
The slight narrowing of the jet occurs in the x–z plane, which is orthogonal to the flapping plane
(figure 15a(i)(iii)). This phenomenon is also observed by other researchers, e.g. Parekh et al. [6]
and Samimy et al. [7]. Parekh et al. [6] attributed this behaviour to an enhanced entrainment in
this plane, thus inducing more ambient fluid to move towards the jet centre. It is worth pointing
out that the streamwise mean velocity profile is different from that in the bifurcating jet, where
two distinct peaks occur in the lateral distribution of the flapping plane (e.g. [48]), that is the
present manipulated jet is different from the bifurcating jet. Figure 15b presents the contours
of time-averaged lateral velocity V ∗ = V/Ue and W ∗ = W/Ue of the excited jet, compared with
unexcited jet. The same lowest contour level is chosen for these plots to facilitate comparison.
Without excitation, the lateral velocity is quite small. Once excited, the V ∗ and W ∗ distributions
change greatly. The contours of V ∗ are positive on the upper side and negative on the lower side
from x∗ ≈ 2.0 onward, suggesting strong spreading occurs in the x–y plane. This is consistent with
the location where the flapping motion occurs, reconfirming that the flapping of the jet causes
spreading. On the other hand, the negative V ∗ on the upper side and the positive on the lower side
ahead of x∗ ≈ 2.0 suggest an entrainment of ambient fluid due to the formation and pairing of the
vortex rings. Similarly, the negative W ∗ on the upper side (z∗ > 0.5) and the positive on the lower
side (z∗ < −0.5) also indicate that ambient fluid is entrained toward the jet. This also validates the
reason why the narrowing stream occurs in the x–z plane, as discussed above. The maximum level
of V ∗ is increased by 600% and that of W ∗ by 300%. In summary, two mechanisms may contribute
to rapid decay of the jet centreline mean velocity. One is due to the enhanced entrainment of the
ambient fluid during early formation and pairing of the vortex rings. The other is the remarkable
spreading associated with the flapping motion in the x–y plane.

7. Conclusion
Investigations have been conducted on the active manipulation of a turbulent round jet
(ReD = 8000) using two unsteady radial minijets separated azimuthally by θ = 60◦ . Two
control parameters are examined, i.e. Cm (1.1%–10.9%) and fe /f0 (0.3–1.24). Measurements were
performed in the (x, y) plane, i.e. the symmetry plane of the two minijets, the (x, z) plane, and a
number of cross-sectional (y, z) planes over x∗ = 0.45–5 using PIV, flow visualization and hotwire
techniques. A comparison is also made with the jet excited by two radial unsteady minijets placed
at θ = 180◦ apart. The following conclusions could be drawn from this work.

1. The jet decay rate K exhibits a strong dependence on Cm and fe /f0 (figure 2). This
behaviour is qualitatively the same as that of a jet under symmetric control (θ =
180◦ ). However, asymmetric excitation increases K by 30%, suggesting a considerably
more effective control strategy. A flapping jet is generated. This flapping behaviour is
responsible for a substantial increase in jet spread, thus giving rise to rapid jet decay.
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

A combination of three factors may have contributed to this observation. Firstly, two
17
minijets, separated by θ = 60◦ , provide an asymmetric excitation, thus generating an
asymmetric flow structure that is more unstable (e.g. [1,18]). Secondly, at a critical

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
Cm such as Cm = 2.0%, vortex rings interact vigorously (e.g. Figure 7c(i)(ii)). Finally,
an excitation at fe /f0 = 1 triggers the formation and alternated pairing of vortex rings.
Henderson [9] suggested that the most essential requirement for mixing enhancement is
to excite a jet-flapping mode. This work provides a new strategy to create a flapping jet,
thus enhancing jet mixing effectively. The excitation may be less effective for higher ReD ,
as reported by e.g. Suzuki et al. [48] and Benard et al. [50].
2. In order to understand the generation of the flapping motion, flow structure under
optimal excitation (Cm = 2.0%, fe /f0 = 1) is analysed in detail. Minijet perturbation gives
rise to early roll-up on the ‘upper’ side because of direct minijet impingement. This
then leads the vortex on the upper side to move inwards to catch up with the adjacent
downstream vortex due to their mutual induction. As a result, the vortex rings are tilted
alternately to the opposite directions, creating a ‘zigzag’ structure and the localized
pairing between the ends of two adjacent rings (figure 12). The vigorous interactions
of these vortical structures entrain ambient fluid into the jet core alternatively from the
upper and the lower side, thus producing a flapping motion.

Data accessibility. This work has no accompanying data.


Authors’ contributions. H.Y. performed the experimental measurements and drafted the paper. Y.Z. proposed the
idea of the study, funded and supervised the project. Y.Z. and R.M.C.S. revised the draft. Y.L. supervised
partially H.Y.’s experimental work at PolyU. All authors regularly discussed the progress during the entire
work.
Competing interests. We declare we have no competing interests.
Funding. Y.Z. wishes to acknowledge support given to him from NSFC through grant 51421063
and from Scientific Research Fund of Shenzhen Government through grants, JCYJ20130329154125496,
KQCX2014052114430139 and KQCX20130627094615415.

References
1. Gutmark EJ, Grinstein FF. 1999 Flow control with noncircular jets. Annu. Rev. Fluid Mech. 31,
239–272. (doi:10.1146/annurev.fluid.31.1.239)
2. Zaman KBMQ. 1999 Spreading characteristics of compressible jets from nozzles of various
geometries. J. Fluid Mech. 383, 197–228. (doi:10.1017/S0022112099003833)
3. Reeder MF, Samimy M. 1996 The evolution of a jet with vortex-generating tabs: real-
time visualization and quantitative measurements. J. Fluid Mech. 311, 73–118. (doi:10.1017/
S0022112096002510)
4. Zaman KBMQ, Hussain AKMF. 1980 Vortex pairing in a circular jet under controlled
excitation. Part 1. General jet response. J. Fluid Mech. 101, 449–491. (doi:10.1017/
S0022112080001760)
5. Wiltse JM, Glezer A. 1993 Manipulation of free shear flows using piezoelectric actuators.
J. Fluid Mech. 249, 261–285. (doi:10.1017/S002211209300117X)
6. Parekh DE, Kibens V, Glezer A, Wiltse JM, Smith DM. 1996 Innovative jet flow control: mixing
enhancement experiments. AIAA Paper No. 96–0308. (doi:10.2514/6.1996-308)
7. Samimy M, Kim JH, Kastner J, Adamovich I, Utkin Y. 2007 Active control of high-
speed and high-Reynolds-number jets using plasma actuators. J. Fluid Mech. 578, 305–330.
(doi:10.1017/S0022112007004867)
8. Smith BL, Glezer A. 2002 Jet vectoring using synthetic jets. J. Fluid Mech. 458, 1–34.
(doi:10.1017/S0022112001007406)
9. Henderson B. 2010 Fifty years of fluidic injection for jet noise reduction. Int. J. Aeroacoustics 9,
91–122. (doi:10.1260/1475-472X.9.1-2.91)
10. Davis MR. 1982 Variable control of jet decay. AIAA J. 20, 606–609. (doi:10.2514/3.7934)
11. Zhou Y, Du C, Mi J, Wang XW. 2012 Turbulent round jet control using two steady minijets.
AIAA J. 50, 736–740. (doi:10.2514/1.J050838)
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

12. New TH, Tay WL. 2006 Effects of cross-stream radial injections on a round jet. J. Turbul. 7,
18
1–20. (doi:10.1080/14685240600847466)
13. Alkislar MB, Krothapalli A, Butler GW. 2007 The effect of streamwise vortices on the

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
aeroacoustics of a Mach 0.9 jet. J. Fluid Mech. 578, 139–169. (doi:10.1017/S0022112007005022)
14. Lardeau S, Lamballais E, Bonnet JP. 2002 Direct numerical simulation of a jet controlled by
fluid injection. J. Turbul. 3, 1–25. (doi:10.1088/1468-5248/3/1/002)
15. Gutmark EJ, Schadow KC, Yu KH. 1995 Mixing enhancement in supersonic free shear flows.
Annu. Rev. Fluid Mech. 27, 375–417. (doi:10.1146/annurev.fl.27.010195.002111)
16. Raman G. 1997 Using controlled unsteady fluid mass addition to enhance jet mixing. AIAA J.
35, 647–656. (doi:10.2514/2.185)
17. Zhang P. 2014 Active control of a turbulent round jet based on unsteady microjets. PhD Thesis,
The Hong Kong Polytechnic University, Hong Kong.
18. Ibrahim MK, Kunimura RR, Nakamura Y. 2002 Mixing enhancement of compressible jets by
using unsteady microjets as actuators. AIAA J. 40, 681–688. (doi:10.2514/2.1700)
19. Longmire EK, Duong LH. 1996 Bifurcating jets generated with stepped and sawtooth nozzles.
Phys. Fluids 8, 978–992. (doi:10.1063/1.868876)
20. Pothos S, Longmire EK. 2001 Asymmetric forcing of a turbulent rectangular jet with a
piezoelectric actuator. Phys. Fluids 13, 1480–1491. (doi:10.1063/1.1357817)
21. Tamburello DA, Amitay M. 2007 Interaction of a free jet with a perpendicular control jet.
J. Turbul. 8, 1–27. (doi:10.1080/14685240601007078)
22. Freund JB, Moin P. 2000 Jet mixing enhancement by high-amplitude fluidic actuation. AIAA
J. 38, 1863–1870. (doi:10.2514/2.839)
23. Zhang P, Cao HL, Zhan J, Zhou Y. 2014 Open-and closed-loop control of a turbulent round
jet based on fluidic means. In Fluid-structure-sound interactions and control, pp. 101–106. Berlin,
Germany: Springer.
24. Mi J, Nobes D, Nathan GJ. 2001 Influence of jet exit conditions on the passive scalar field of
an axisymmetric free jet. J. Fluid Mech. 432, 91–125.
25. Gutmark E, Ho CM. 1983 Preferred modes and the spreading rates of jets. Phys. Fluids 26,
2932–2938. (doi:10.1063/1.864058)
26. Raman G, Cornelius D. 1995 Jet mixing control using excitation from miniature oscillating jets.
AIAA J. 33, 365–368. (doi:10.2514/3.12444)
27. M’Closkey RT, King JM, Cortelezzi L, Karagozian AR. 2002 The actively controlled jet in
crossflow. J. Fluid Mech. 452, 325–335. (doi:10.1017/S0022112001006589)
28. Choi JJ, Annaswamy AM, Lou H, Alvi FS. 2006 Active control of supersonic impingement
tones using steady and pulsed microjets. Exp. Fluids 41, 841–855. (doi:10.1007/s00348-006-
0189-7)
29. Ho CM, Huang LS. 1982 Subharmonics and vortex merging in mixing layers. J. Fluid Mech.
119, 443–473. (doi:10.1017/S0022112082001438)
30. Hussain AKMF. 1983 Coherent structures-reality and myth. Phys. Fluids 26, 2816–2850.
(doi:10.1063/1.864048)
31. Oberleithner K, Paschereit CO, Wygnanski I. 2014 On the impact of swirl on the growth of
coherent structures. J. Fluid Mech. 741, 156–199. (doi:10.1017/jfm.2013.669)
32. Ho CM, Huerre P. 1984 Perturbed free shear layers. Annu. Rev. Fluid Mech. 16, 365–424.
(doi:10.1146/annurev.fl.16.010184.002053)
33. Zaman KBMQ, Hussain AKMF. 1981 Turbulence suppression in free shear flows by controlled
excitation. J. Fluid Mech. 103, 133–159. (doi:10.1017/S0022112081001274)
34. Goldschmidt VW, Bradshaw P. 1973 Flapping of a plane jet. Phys. Fluids 16, 354–355.
(doi:10.1063/1.1694348)
35. Michalke A, Hermann G. 1982 On the inviscid instability of a circular jet with external flow.
J. Fluid Mech. 114, 343–359. (doi:10.1017/S0022112082000196)
36. Cohen J, Wygnanski I. 1987 The evolution of instabilities in the axisymmetric jet.
Part 1. The linear growth of disturbances near the nozzle. J. Fluid Mech. 176, 191–219.
(doi:10.1017/S0022112087000624)
37. Corke TC, Kusek SM. 1993 Resonance in axisymmetric jets with controlled helical-mode input.
J. Fluid Mech. 249, 307–336. (doi:10.1017/S0022112093001193)
38. Webster DR, Longmire EK. 1997 Vortex dynamics in jets from inclined nozzles. Phys. Fluids 9,
655–666. (doi:10.1063/1.869223)
39. Iio S, Hirashita K, Katayama Y, Haneda Y, Ikeda T, Uchiyama T. 2013 Jet flapping control with
acoustic excitation. J. Flow Control Meas. Vis. 1, 49–56. (doi:10.4236/jfcmv.2013.12007)
Downloaded from http://rspa.royalsocietypublishing.org/ on July 27, 2016

40. Silva CB, Metais O. 2002 Vortex control of bifurcating jets: a numerical study. Phys. Fluids 14,
19
3798–3819. (doi:10.1063/1.1506922)
41. Reynolds WC, Parekh DE, Juvet PJD, Lee MJD. 2003 Bifurcating and blooming jets. Annu. Rev.

rspa.royalsocietypublishing.org Proc. R. Soc. A 472: 20160417


...................................................
Fluid Mech. 35, 295–315. (doi:10.1146/annurev.fluid.35.101101.161128)
42. Liepmann D, Gharib M. 1992 The role of streamwise vorticity in the near-field entrainment of
round jets. J. Fluid Mech. 245, 643–668. (doi:10.1017/S0022112092000612)
43. Lasheras JC, Cho JS, Maxworthy T. 1986 On the origin and evolution of steamwise
vortical structures in a plane, free shear layer. J. Fluid Mech. 172, 231–258. (doi:10.1017/
S0022112086001726)
44. Bernal LP, Roshko A. 1986 Streamwise vortex structure in plane mixing layers. J. Fluid Mech.
170, 499–525. (doi:10.1017/S002211208600099X)
45. Mi J, Nathan GJ, Luxton RE. 2001 Mixing characteristics of a flapping jet from a self-exciting
nozzle. Flow Turbul. Combustion 67, 1–23. (doi:10.1023/A:1013544019463)
46. Terashima O, Sakai Y, Nagata K, Ito Y, Onishi K, Shouji Y. 2015 On the velocity and pressure
statistics during the flapping motion in a planar turbulent jet. Int. J. Heat Fluid Flow 53, 56–67.
(doi:10.1016/j.ijheatfluidflow.2015.01.007)
47. Chiekh MB, Ferchichi M, Bera JC. 2011 Modified flapping jet for increased jet spreading using
synthetic jets. Int. J. Heat Fluid Flow 32, 865–875. (doi:10.1016/j.ijheatfluidflow.2011.06.004)
48. Suzuki H, Kasagi N, Suzuki Y. 2004 Active control of an axisymmetric jet with distributed
electromagnetic flap actuators. Exp. Fluids 6, 498–509. (doi:10.1007/s00348-003-0756-0)
49. Parekh DE, Reynolds WC, Mungal MG. 1987 Bifurcation of round air jets by dual-mode
acoustic excitation. AIAA Paper No. 87–0164. (doi:10.2514/6.1987-164)
50. Benard N, Balcon N, Touchard G, Moreau E. 2008 Control of diffuser jet flow: turbulent kinetic
energy and jet spreading enhancements assisted by a non-thermal plasma discharge. Exp.
Fluids 45, 333–355. (doi:10.1007/s00348-008-0483-7)

View publication stats

You might also like