Sunteți pe pagina 1din 13

RESEARCH ARTI CLE

Measurement of ambient uid entrainment during laminar vortex


ring formation
Ali B. Olcay Paul S. Krueger
Received: 11 December 2006 / Revised: 14 September 2007 / Accepted: 14 September 2007 / Published online: 14 October 2007
Springer-Verlag 2007
Abstract Planar laser induced uorescence (PLIF) and
digital particle image velocimetry (DPIV) combined with
Lagrangian coherent structure (LCS) techniques are uti-
lized to measure ambient uid entrainment during laminar
vortex ring formation and relate it to the total entrained
volume after formation is complete. Vortex rings are
generated mechanically with a piston-cylinder mechanism
for a jet Reynolds number of 1,000, stroke ratios of 0.5, 1.0
and 2.0, and three velocity programs (Trapezoidal, trian-
gular negative and positive sloping velocity programs). The
quantitative observations of PLIF agree with both the total
ring volume and entrainment rate measurements obtained
from the DPIV/LCS hybrid method for the jet Reynolds
number of 1,000, trapezoidal velocity program and stroke
ratio of 2.0 case. In addition to increased entrainment at
smaller stroke ratios observed by others, the PLIF results
also show that a velocity program utilizing rapid jet initi-
ation and termination enhances ambient uid entrainment.
The observed trends in entrainment rate and nal entrained
uid fraction are explained in terms of the vortex roll-up
process during vortex ring formation.
1 Introduction
Transient ejection of a jet from a nozzle is a common ow
conguration which engenders the formation of a vortex
ring. During the jet ejection, the shear layer which sepa-
rates at the nozzle lip rolls up and entrains some of the
ambient uid into the forming vortex ring as described by
Didden (1979). Consequently, both ejected uid which
comes from inside the cylinder and ambient uid which is
pulled from the vicinity of the nozzle must be accelerated
as the ring forms. The convective nature of the entrainment
process is directly relevant for a wide variety of problems
including cooling of a CPU unit (Kercher et al. 2003),
extinguishing oil well res at places where bringing man-
power and technology can be very expensive (Akhmetov
et al. 1980), mixing two different uids, and transferring
mass from one location to another. Hence, a detailed
understanding of the entrainment mechanics in transient
jets could be applied to enhance or lessen entrainment
effects in a variety of applications.
Most work on entrainment in vortex rings to date has
considered formed or steady vortex rings. In this state, a
closed streamline encircles the vortex ring in a frame of
reference moving at the vortex ring velocity. Thus, the uid
transported within the ring in the vortex bubble is clearly
dened as described by Shariff and Leonard (1992).
Maxworthy (1972) conducted experiments using dye
visualization and hydrogen bubble techniques to study the
diffusion of vorticity, which results in entrainment after the
vortex bubble is formed. He concluded that while some of
the vorticity diffuses by pulling irrotational uid inside the
vortex bubble making vortex bubble larger in size, some
stays behind the traveling vortex ring forming a wake.
Maxworthy (1977) also performed some experiments to
study the vortex ring formation as well as evolution of the
A. B. Olcay
Department of General Engineering,
University of Wisconsin-Platteville,
1 University Plaza, Platteville, WI 53818, USA
e-mail: olcaya@uwplatt.edu
P. S. Krueger (&)
Department of Mechanical Engineering,
Southern Methodist University, P.O. Box 750337,
Dallas, TX 75275, USA
e-mail: pkrueger@engr.smu.edu
1 3
Exp Fluids (2008) 44:235247
DOI 10.1007/s00348-007-0397-9
vortex rings. He noted the effect of Reynolds number on
the formation process, but he did not comment about its
effect on entrainment. Fabris and Liepmann (1997) ana-
lyzed vortex ring formation, and they concluded that there
are three distinct regions in a steady vortex ring, namely
the core region where rotational ow is present, an inter-
mediate region where irrotational ow exists in the form of
ejected and entrained uid, and nally an external region
where potential ow encloses the moving vortex bubble.
Dabiri and Gharib (2004) investigated entrainment using
steady bulk counter-ow to hold the rings in the eld of
view and streamlines obtained from DPIV to identify the
bubble volume. They observed the rings contained up to
65% entrained uid volume when they were completely
formed. It was also observed that entrainment fraction
could be increased by using a smaller stroke ratio. A dif-
fusive uid entrainment model was developed by relating
the ratio of entrained uid ux (i.e., time rate of change of
entrained uid volume) to the total uid ux in the dissi-
pation region behind the formed ring with the ratio of
ambient uid energy loss rate by viscous dissipation to
ambient uid energy entering the dissipation region. These
ratios where expressed in terms of the vortex rings gov-
erning parameters (e.g., velocity, volume of the vortex ring,
and diameter of vortex ring generator).
For formed rings experiencing periodic forcing, the
unstable manifold of the forward stagnation point no longer
coincides with the stable manifold of the rear stagnation
point and intersections between these manifolds identify
lobes of uid which can cross the ring boundary
(entrainment or detrainment) as the ow evolves. Shariff
et al. (2006) studied this process for numerically generated,
time-periodic vortex rings using dynamical systems theory.
By monitoring evolution of the lobes and changing oscil-
lation amplitude of the periodic disturbance, they showed
that the exchanged volume can be increased when a higher
oscillation amplitude is used. They also noted the quanti-
tative similarity between their results and detrainment from
experimentally generated turbulent rings.
Shadden et al. (2006) studied empirically generated
vortex rings and observed entrainment and detrainment by
lobe dynamics for nominally steady (i.e., quasi-steady),
aperiodic ows. In this case, the stable and unstable man-
ifolds delineating the vortex boundary were identied as
ridges in the nite-time Lyapunov exponent (FTLE) eld
obtained from digital particle image velocimetry (DPIV)
data of the velocity eld. They also computed the bubble
volume identied by the ridges [called Lagrangian coher-
ent structures (LCSs)] and compared it with that
determined from streamlines.
Although all of these studies highlight various features
of entrainment during the steady (or quasi-steady) phase of
laminar vortex ring motion when a closed vortex bubble
has formed, none of them address the process of entrain-
ment during initial ring formation and roll-up. Yet, the bulk
of the entrained uid in a steady vortex ring is acquired
during the formation process (Dabiri and Gharib 2004).
Indeed, Auerbach (1991) made the distinction between
convective entrainment during shear layer roll-up phase
and diffusive entrainment after the vortex ring is formed.
While he was unable to provide quantitative measure of
ambient uid entrainment during vortex ring formation, he
concluded that depending on the formation details as much
as 40% of the uid carried with a steady ring can be
ambient uid.
The objective of this study is to measure ambient uid
entrainment during laminar vortex ring formation, and
evaluate the effect of vortex ring formation parameters on
the entrainment process. Since there is no closed volume
associated with the vortex during the formation process
(i.e., prior to achieving a nearly constant translational
velocity), we instead focus on the rate at which uid is
entrained into the forming vortex spiral. Explicit identi-
cation of the entrance to the spiral will be given in
Sect. 3.3. The experimental observations are made using
planar laser induced uorescence (PLIF) and DPIV. Using
dye as a Lagrangian marker of the ejected uid, PLIF
allows direct observation of the vortex spiral during the
formation process. DPIV combined with LCS techniques
gives analogous results, but also includes velocity infor-
mation allowing a more detailed analysis. Both data sets
(PLIF and DPIV) can be used to deduce the size of the
vortex bubble once formed, indicating the overall entrain-
ment. Using these techniques, the present investigation
studies the evolution of entrainment during ring formation
under the inuence of different jet velocity programs and
ejected jet length-to-diameter ratios (L/D).
2 Experimental setup and techniques
A schematic of the experimental apparatus is given in
Fig. 1. The experimental apparatus consisted of a piston
cylinder mechanism for generating the vortex rings, a water
tank, and a pressurized tank to drive the piston. The water
tank was 61 cm wide, 61 cm deep and 244 cm long. The
walls of the tank were 1.27 cm thick glass for ow visu-
alization purposes. The piston and cylinder of the vortex
ring generator were made from high-impact strength PVC
rod and clear PVC schedule-40 pipe, respectively. The
cylinders inner diameter (D) was 3.73 cm. A critical
parameter for vortex ring formation was the length-to-
diameter ratio (L/D), dened as the ratio of the total piston
displacement (during jet ejection) to the piston diameter.
The outer surface of cylinder nozzle was machined to have
a wedge with a tip angle (a) of 7 to ensure clean ow
236 Exp Fluids (2008) 44:235247
1 3
separation at the nozzle exit plane. The pistoncylinder
mechanism was connected to a 30-gallon tank which was
pressurized to 15 psig (103 kPa gage) to actuate the piston
during measurements. A proportional solenoid valve
(SD8202, ASCO Valve Inc.) and an inline ultrasonic ow
rate probe (ME19PXN, Transonic Systems Inc.) were used
to control and measure the volumetric ow rate, respec-
tively. The piston velocity was determined from the ratio of
volumetric ow rate to the piston area. An in-house-code
programmed in Labview (National Instruments) provided
feedback control of the ow rate allowing the piston to
follow an arbitrary velocity program.
To generate the vortex rings, the piston was commanded
to execute nite duration jet pulses. Three different jet
velocity programs were considered: trapezoidal, triangular
positive sloping (PS), and triangular negative sloping (NS).
Examples of all three are shown in Fig. 2. Jet Reynolds
number (Re
J
) is calculated based on the pistons maximum
velocity (U
M
), piston diameter (D) and the uid viscosity
(m), namely,
Re
J

U
M
D
m
: 1
In Fig. 2 solid lines show the commanded velocity, and
hollow triangles indicate the measured piston velocity. As
seen from the plots, there is a very good agreement
between commanded and measured velocities. Repeat-
ability of velocity programs was better than 10% in Re
J
and
5% in L/D. While acceleration and deceleration periods
were chosen to last 0.1t
p
(t
p
being the pulse duration) for a
trapezoidal velocity program, triangular PS and triangular
NS velocity programs commanded the piston with 0.9t
p
and 0.1t
p
in the acceleration phase, 0.1t
p
and 0.9t
p
during
the deceleration phase, respectively. The triangular PS and
NS cases introduce the effects of non-impulsive jet initia-
tion and termination, respectively.
In order to study the effects of piston velocity program
and L/D ratio, a number of experiments were performed. A
summary of tests used in this study along with the utilized
techniques are given in Table 1.
For PLIF measurements a cover plate similar to that
used by Johari (1995) was initially placed over the end of
the cylinder (see Fig. 1). A hole in the plate allowed dyed
uid to be drawn into the cylinder while the plate was in
place ensuring the uid in the cylinder and the ambient
uid were initially separate. Just before the test began the
cover plate was drawn up vertically. Velocity of the cover
plate was 0.49 0.10 cm/s. This velocity was small
enough to cause minimal dye disturbance prior to tests.
When the plate was clear, the piston was actuated. Using
this procedure, only dyed uid was in the cylinder when the
jet was initiated and therefore the dye may be considered as
a Lagrangian marker for the ejected uid. Fluorescein dye
at 4.7 10
7
M was used as the uid marker. The Schmidt
number for uorescein was about 1,000 as suggested by
Green (1995), so the dye tracked the uid motion, but not
vorticity diffusion.
An Argon ion laser (Innova 70-2, Coherent Inc.) was
used to illuminate the dye. A 0.15 cm thick laser sheet was
obtained using a cylindrical lens. A black and white 8-bit
digital CCD camera (UP-1830, Uniq Vision Inc.) was
placed perpendicular to the laser sheet to record the ow
evolution at 30 Hz. The recorded eld had a 1,024
1,024 pixels spatial resolution with a 4.65D 4.65D
eld of view.
Digital particle image velocimetry (DPIV) was used to
obtain velocity eld data for the vortex ring formation
process. The DPIV system consisted of a pair of frequency
doubled, pulsed Nd:YAG lasers (Vlite200, LABest Inc.),

Pressure
regulator
Pressurized
tank
Proportional
solenoid valve
Flowrate meter
Laser sheet
Cover plate
and lifting
mechanism
Vortex ring
generator
Water tank
Pressurized
air
Feedback control
through Labview
Water
Piston
U
p
(t)
D
Fig. 1 Schematic of
experimental apparatus
Exp Fluids (2008) 44:235247 237
1 3
optics to transform the laser beam into a 0.1 cm thick laser
sheet, and a delay generator (555 Pulse Delay Generator,
Berkeley Nucleonics Corporation) for synchronizing the
laser and the camera. The particles used to seed the ow
were 1520 lm diameter neutrally buoyant silver-coated
hollow glass spheres (SH400S20, Potters Industries Inc.).
The obtained 1,024 1,024 pixels particle images were
recorded, paired and processed using Pixel Flow (FG
Group LLC.), which uses a cross-correlation algorithm
similar to the one described by Willert and Gharib (1991).
A laser pulse separation of 22.7 ms was used, giving
maximum particle displacements of 78 pixels for a
3.0D 3.0D eld of view. With 32 32 pixel interroga-
tion windows at 50% (16 pixels) overlap, the spatial
resolution of the resulting vector elds was 0.094D
0.094D. To improve the accuracy, the data were processed
a second time with a window-shifting algorithm as
described by Westerwheel et al. (1997). The uncertainty of
velocity measurements was 0.04 pixels as stated by
Westerwheel et al. (1997), which was less than 1% for the
majority of the measured velocity eld.
The DPIV data were also used to obtain Lagrangian
information about the ow as expressed using the nite
time Lyapunov exponent (FTLE). Details of this approach
may be found in Shadden et al. (2005); however, a brief
overview will be presented here for completeness. The
equation describing the trajectory of a uid particle at
position x
0
at time t
0
may be expressed as
_ xt; t
0
; x
0
Vxt; t
0
; x
0
; t 2
where xt
0
; t
0
; x
0
x
0
:
The right side of Eq. (2) can be attained from the DPIV
velocity eld data. The solution to (2) is a ow map
/
t
0
T
t
0
x
0
describing the position at time t = t
0
+ T of the
uid particle initially at x
0
at time t
0
namely;
/
t
0
T
t
0
x
0
xt
0
T; t
0
; x
0
: 3
Then the nite time Lyapunov exponent (FTLE) is dened as
t/t
p
U
p
(
t
)
/
U
M
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
1.2
Commanded velocity
Measured velocity
Commanded velocity
Measured velocity
(a)
t/t
p
U
p
(
t
)
/
U
M
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
1.2 (b)
t/t
p
U
p
(
t
)
/
U
M
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
1.2
Commanded velocity
Measured velocity
(c)
Fig. 2 Typical piston velocity
programs for Re
J
= 1,000,
L/D = 2.0. a Trapezoidal
velocity program (t
P
= 2.77 s),
b triangular positive sloping
(t
P
= 4.98 s), and c triangular
negative sloping (t
P
= 4.98 s)
velocity programs
Table 1 Table of the tested cases
L/D Re
J
Velocity program Flow analysis technique
0.5, 1.0 1,000 Trapezoidal PLIF
2.0 1,000 Trapezoidal PLIF, Streamline, and LCS
0.5, 1.0 1,000 Triangular NS PLIF and LCS
2.0 1,000 Triangular NS PLIF and LCS
0.5, 1.0 1,000 Triangular PS PLIF
2.0 1,000 Triangular PS PLIF and LCS
238 Exp Fluids (2008) 44:235247
1 3
r
T
t
0
x
1
T j j
ln

k
max
p
4
where k
max
is the maximum eigenvalue of
r/
t
0
T
t
0
x

r/
t
0
T
t
0
x

5
and ()* denotes the adjoint operation. It can be shown
(Shadden 2005) that the separation of particles advected by
the ow is proportional to e
r
T
t
0
x T j j
to highest order. Hence,
the FTLE is roughly a measure of the maximum expansion
rate of particle pairs advected by the ow.
Lagrangian coherent structures (LCS) are dened as the
ridges in the FTLE eld. Shadden et al. (2005) show that
the ux across a LCS scales like
1
T j j
and thus, for large |T| a
LCS can be treated as a material line or transport barrier in
the ow. Additionally, LCS obtained by forward (T [0)
and backward (T \0) time integration recovers the stable
(also called repelling LCS) and unstable manifolds (also
called attracting LCS), respectively, surrounding a vortex
ring. Since LCSs behave like material lines, they can
identify the vortex bubble boundaries. This was illustrated
by Shadden et al. (2006), who combined the attracting and
repelling LCS to study entrainment of a formed ring. Since
the formulation applies equally well to unsteady ows, it
can be used to identify the vortex boundary during ring
formation and hence, is a useful tool for studying
entrainment during this phase of ring evolution as well.
LCSs in this study were calculated by using a software
package called ManGen developed by Francois Lekien and
Chad Coulliette in 2001 (http://www.lekien.com/*
francois/software/mangen/). ManGen provided the FTLE
eld by computing Eq. (4) for a grid of massless particles
placed in the domain and advected using the given velocity
eld. For the present investigation, computation of r was
performed with a uniform grid of 0.0067D resolution to
produce clear ridges for both attracting and repelling LCS.
3 Results
3.1 Qualitative observations
Figure 3 illustrates vortex ring formation from a trapezoi-
dal, triangular NS and PS velocity programs for Re
J
of
1,000 and L/D = 2.0. All the rings in Fig. 3 travel from left
to right, and t* in the gures is dened as t

t
p
: In these
gures, gray pixels represent the ejected uid coming from
inside the cylinder, and black pixels represent ambient uid
which was initially outside the nozzle. During jet ejection
(0 t* 1), entrainment is apparent through the growing
black spiral, but the volume of entrained uid is clearly
much less than the ejected uid.
Comparison of the spiral formation among the velocity
programs shows distinct differences. Since trapezoidal and
triangular NS velocity programs have a similar start up
acceleration, both produce a long tightly wound spiral
during the initial jet ejection. In contrast, the initial spiral
for the PS case is not wound tightly because the slow jet
initiation provides much less vorticity initially yielding
fewer spiral loops, and the width of each loop is larger. In
general, trapezoidal and triangular NS velocity programs
have steeper start up accelerations than the triangular PS
velocity programs. This steep start up acceleration is
responsible for high velocity gradients at the inner nozzle
wall, which produce stronger vorticity at start up. The

y
t
i
c
o
l
e
V

l
a
d
i
o
z
e
p
a
r
T
p
o
r
g
m
a
r



)
S
N
(

e
l
g
n
a
i
r
T
t
i
c
o
l
e
V
y
o
r
P

g
m
a
r



)
S
P
(

e
l
g
n
a
i
r
T
t
i
c
o
l
e
V
y
o
r
P

g
m
a
r

t
*
=0.5 t
*
=1.0 t
*
=1.5 t
*
=2.0
t
*
=0.5 t
*
=1.0 t
*
=1.5 t
*
=2.0
t
*
=0.5 t
*
=1.0 t
*
=1.5 t
*
=2.0
Fig. 3 PLIF ow visualization
vortex rings generated by
trapezoidal, triangular NS and
PS velocity programs for Re
J
=
1,000 and L/D = 2.0. Gray and
black pixels represent the
ejected and ambient uid,
respectively
Exp Fluids (2008) 44:235247 239
1 3
result is strong Biot-Savart induction, a tightly wound
spiral and, as will be shown later, high ambient uid
entrainment. Additionally, while the vortex rings obtained
with trapezoidal and triangular NS velocity programs leave
behind a noticeable quantity of ejected uid, rings gener-
ated by triangular PS velocity program pull in nearly all the
ejected uid. This results in proportionally more ambient
uid within the ring for the former case. Finally, it is noted
that most of the entrainment occurs after the piston has
stopped (see Fig. 3 for t*[1.0). Once the piston stops, the
ejected uid boundary is no longer held out by the jet and it
may contract under the inuence of the ring vorticity,
making a larger area available for entrainment of ambient
uid. The vorticity created during shear layer roll up drives
the ambient uid entrainment not only during formation
phase but also after piston has stopped. These observations
agree with Diddens (1979) results.
Qualitative observations of the ring formation are also
obtained using DPIV by computing the streamfunction in a
frame of reference moving with the vortex ring. The
velocity eld is converted to this reference frame by sub-
tracting the ring velocity based on the position of the peak
ring vorticity. To obtain accurate ring velocity measure-
ments, a Gaussian t of the vortex core is used to obtain
subgrid estimates of vortex location and a third order
polynomial t of the results is used for computing velocity.
Then, Stokes streamfunction is obtained by solving the
governing equation
1
r
o
2
W
ox
2

o
or
1
r
oW
or

x
h
6
with a second order accurate nite difference method using
the vorticity eld obtained from DPIV data and the velocity
data as boundary conditions.
Figure 4 displays DPIV measurements of vortex ring
evolution for Re
J
= 1,000 and L/D = 2.0. The dashed line in
Fig. 4 represents the stagnation streamline which identies
the boundary of the vortex bubble. Figure 4a shows that, in
the eld of view, the W = 0 streamline does not reach r = 0
on the backside of the ring while the jet is on. Although
data was collected only for x [0, we can conclude that the
W = 0 streamline does not reach r = 0 for x \0 while the
jet is on since instantaneous streamlines cannot intersect.
Thus, mass is still entering the ring in Fig. 4a. The W = 0
streamline may, however, close on the outer annulus of the
nozzle. Once the piston has stopped moving, a stagnation
point develops behind the ring, forming a closed bubble as
shown in Fig. 4b, c.
The vorticity distribution is also given in Fig. 4. During
jet ejection the vortex core is small and moves slightly
outward, above the piston radius (i.e., 0.5D). As the ring
continues to form (Fig. 4b), the bounding streamline
obtains fore-aft symmetry. During further evolution of the
vortex ring, vorticity starts to diffuse out of the vortex
bubble (Fig. 4c) causing enlargement of the vortex bubble
volume as mentioned by Maxworthy (1972). The move-
ment of the vorticity core during ring formation determines
the area where induced ambient uid entrainment takes
place.
Once the DPIV velocity vector elds are obtained from
experiments, FTLE elds are calculated using ManGen
(see Fig. 5a). It is noted that high FTLE values (i.e., ridges)
illustrate the LCS (also called attracting LCS since T\0)
as stated by Shadden et al. (2006). During post-processing,
a threshold is applied to FTLE elds to identify the LCS
and locate Dr to be used for volume and entrainment cal-
culations, respectively (see Fig. 5b). This is discussed
further in Sect. 3.3.
The time evolution of the LCS for a trapezoidal velocity
program for Re
J
= 1,000 and L/D = 2.0 is shown in Fig. 6.
The repelling and attracting LCSs in Fig. 6ac were
obtained with forward integration (i.e., T [0) and back-
ward integration (i.e., T \0), respectively. Combining
these repelling and attracting LCSs generate a closed
transport barrier dening the vortex bubble as described by
Shadden et al. (2006). The extended back side can be
observed in Fig. 6a, b representing the mass that will
comprise the nal vortex bubble. This unique property of
LCS determines the volume associated with the vortex ring
x / D
0 0.5 1 1.5 2 2.5
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(a)
x / D
0 0.5 1 1.5 2 2.5
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(b)
x / D
r

/

D
r

/

D
r

/

D
0 0.5 1 1.5 2 2.5
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(c)
Fig. 4 Contour plots of vorticity with the stagnation streamline indicated by dashed lines. ac are at t* = 0.72, 1.08, and 1.81, respectively
240 Exp Fluids (2008) 44:235247
1 3
before it is completely formed. It is also noted that the LCS
in Fig. 6b delineates the primary vortex and the stopping
vortex as observed by Didden (1979). Since the stopping
vortex does not enter into the forming ring, the LCS
veries this ow is outside of the vortex bubble. Lastly,
Fig. 6c illustrates the LCS after the ring is completely
formed. The LCS representing boundary of the vortex
bubble demonstrates near fore-aft symmetry once ring
formation is complete (in agreement with Fig. 4c).
3.2 Entrainment for a trapezoidal velocity program
For a quantitative assessment of the entrainment, we rst
focus on the trapezoidal velocity program as a canonical
case frequently studied in the literature. Figure 7 illustrates
PLIF of a vortex ring just after formation is complete for
Re
J
= 1,000 and L/D = 2.0. To nd the boundary of the
vortex bubble oX from such data, an edge detection
algorithm was applied to the images after thresholding.
While the front edge of the moving vortex bubble can be
clearly observed with the PLIF technique, the rear edge of
the vortex bubble cannot be identied by this technique.
Therefore, front-back symmetry is assumed using the upper
and lower boundary of the vortex bubble to dene the
mirror axis as indicated in Fig. 7. As noted earlier, this
assumption applies well for a formed ring (Fig. 4b, c).
Integrating over the hatched region in Fig. 7 and assuming
axisymmetry gives the vortex bubble volume (V
B
).
Although the uncertainty of the volume calculation
depends on the location of mirror axis, threshold value for
edge detection, and image quality, uncertainty analysis
shows that overall uncertainty in V
B
falls below 7%.
Once the vortex bubble boundary is determined, the
volume of ejected uid within this volume (V
EJ
) is calcu-
lated by integrating the volume of the gray pixels within
oX (again assuming axisymmetry). This approach accu-
rately obtains the amount of ejected uid which remains in
the vortex bubble at the end of a piston stroke since it does
not consider ejected uid not entrained into the vortex
Fig. 5 a shows the color contour plots of FTLE eld at |T| = 5.01 s(|T|/t
p
= 1.81); b illustrates the LCS after thresholding. It is noted that ambient
uid entrainment occurs through Dr
x / D
r

/

D
0 0.5 1 1.5 2 2.5
x / D
0 0.5 1 1.5 2 2.5
x / D
0 0.5 1 1.5 2 2.5
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
r

/

D
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
r

/

D
-1.25
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
1.25
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(a)
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(b)
(1/s)
8
7
6
5
4
3
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8
(c)
Primary
vortex
Stopping vortex
Fig. 6 Vortex evolution observed from LCS. The solid line is the repelling LCS and the dashed line is the attracting LCS. ac illustrates
evolution of the vortex bubble at t* of 0.72 (T = 1.99 s), 1.08 (T = 2.99 s) and 1.81 (T = 5.01 s), respectively
Exp Fluids (2008) 44:235247 241
1 3
bubble (see PLIF data for trapezoidal and triangular NS
when t* 1.5 in Fig. 3). Therefore, once the ejected uid
in the vortex bubble is known, volume of entrained
ambient uid (V
E
) can be calculated as
V
E
t V
B
t V
EJ
t: 7
Dye diffusion in the vortex core, however, leads to ambi-
guity in identication of ejected versus entrained uid in
the core. The related uncertainty in V
EJ
is less than 4.2%
for t* 0.5.
The preceding analysis employing fore-aft symmetry to
identify the boundary of a formed ring may be readily
extended to a forming ring. During jet ejection, fore-aft
symmetry is clearly violated near the centerline because
there is no stagnation point on the back side of the ring, but
near the forming spiral such symmetry holds approxi-
mately. This is illustrated in Fig. 8 where the boundary
obtained from the front of the ring is reected about the
mirror axis to compare with the back side of the ring. The
lack of symmetry on the back side near the axis is of no
consequence for computing V
E
, however, since the ejected
uid in this region is subtracted out. Thus, V
E
(t) computed
with this procedure, in essence, identies the uid
entrained into the spiral across the arc AB in Fig. 8. Spe-
cically,
dV
E
dt
computed from these results should give an
accurate measurement of the rate at which ambient uid is
entrained into the ring during formation. This statement
will be justied experimentally in Sect. 3.3.
Results of the PLIF volume calculations for Re
J
= 1,000,
L/D = 2.0, and a trapezoidal velocity program are shown in
Fig. 9. Volume by PLIF refers the volume calculation
performed on hatched region given in Fig. 7, and ejected
volume in the ring is the volume of gray pixels (i.e., ejected
uid) in the hatched region of Fig. 7. The results show that
the computed V
B
and V
EJ
increase at nearly identical rates
during formation so that V
E
is small during this phase.
Once the piston stops, ejected uid entry slows dramati-
cally since any ejected uid left at the vicinity of the nozzle
does not have enough momentum to catch the vortex
bubble. Nevertheless, the vortex bubble volume continues
to increase after the jet stops until it reaches an asymptotic
value of V
B
/V
EJ
= 1.25. This nal increase in the volume is
mostly due to the ambient uid entrainment. Therefore, the
vortex ring velocity (see Fig. 10) decelerates in this region
(between t* of 1.0 and 1.5) since momentum initially
supplied by the piston needs to be shared with this addi-
tional mass, namely the entrained ambient uid. This
indicates most of the ambient uid is entrained during the
impulse-preserving phase of motion after the jet stops (as
observed qualitatively in Sect. 3.1).
The PLIF data shows a nearly constant bubble volume
after ring formation is complete, verifying that the ring is
formed since its shape and volume remain unchanged. In
reality, the vortex bubble continues to increase slowly due
Fig. 7 The components of a moving vortex bubble obtained by PLIF
for Re
J
= 1,000 and L/D = 2.0
Mirror
axis
A
B
Fig. 8 Illustration of the rear edge obtained in PLIF images assuming
fore-aft symmetry
t / t
P
V
B
/
V
E
J
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
Volume by PLIF
Volume by streamlines
Volume by LCS
Ejected Volumeinring (PLIF)
Fig. 9 Vortex bubble volume calculation using PLIF, DPIV and LCS
data (Re
J
= 1,000, L/D = 2.0)
242 Exp Fluids (2008) 44:235247
1 3
to the vorticity diffusion as mentioned by Dabiri and
Gharib (2004) and Maxworthy (1972). This is not apparent
in the PLIF data due to slow dye diffusion (high Schmidt
number). As a consequence, the PLIF data indicates vol-
ume obtained during formation only and not as a result of
subsequent vorticity diffusion. The streamfunction volume
calculation is also given in Fig. 9 and veries the volume
increase due to vorticity diffusion. Streamfunction volume
calculations are only obtained after a vortex bubble is
formed so that a closed streamline (see Fig. 4b, c) is
obtained.
Using the stagnation streamline dening the vortex
bubble as shown in Fig. 4, V
B
can also be computed once
the ring is formed. As before, axisymmetry is assumed in
the volume calculation. The major source of uncertainty
comes from the vortex ring velocity calculation and is less
than 12%. The streamfuction volume calculation agrees
with the PLIF results to within the experimental error just
after ring formation is complete (1.35 t* 2.0). As time
proceeds, the vortex bubble volume obtained from the
streamfunction starts to increase, reecting the effect of
vorticity diffusion as the ring advects downstream. Since
vortex bubble volume increases by this additional mass, we
can see that vortex ring slows down after t* of 1.75 (see
Fig. 10).
The lled diamonds in Fig. 9 represent the vortex bub-
ble volume calculation from the LCS technique. These
values were obtained by integrating the volume inside the
LCS boundaries (see Fig. 6) assuming axisymmetry. The
uncertainty in these volume calculations comes mainly
from threshold values applied for repelling and attracting
LCSs and is less than 3%. The LCS data indicate a constant
bubble volume for 0.7 t* 1.4, even though the PLIF
data indicate the ring is not formed until t* [1.5. This is
because the nature of the LCS as transport barriers allows
them to identify the uid volume eventually to appear in
the ellipsoidal volume of the formed vortex ring bubble,
even before the bubble is formed (see Fig. 6a, b). The LCS
data for t* [1.4 illustrates an increase in vortex bubble
volume in agreement with the streamfunction volume
calculation to within experimental uncertainty.
As with the streamfunction data, the LCS data agree
with the volume obtained from the PLIF results to within
experimental uncertainty, conrming the validity of the
fore-aft symmetry assumption used in computing the
bubble volume from the PLIF images, at least in the case of
a completely formed ring.
3.3 Entrainment rate
The rate of uid entrainment into the vortex ring (Q
E
) is
dened as
Q
E

dV
E
dt
8
where V
E
is the volume of ambient uid in the vortex ring
spiral. This can be estimated from the PLIF data, but the
PLIF data relies on the assumption of front-back symmetry
during ring formation. The LCS data, on the other hand,
can obtain Q
E
directly. Specically, we consider the vol-
ume ow rate into the entrance gap of the vortex spiral as
identied by the attracting LCS. For convenience, we
consider the gap Dr = r
o
r
i
identied along the line
connecting vortex cores as shown in Fig. 11a. This is
analogous to entrance of the vortex spiral identied in the
PLIF data as can be seen by comparing Fig. 11a, b.
Although taken from different runs, the location and the
magnitude of Dr for the LCS and PLIF data show rea-
sonable agreement. In particular, the entrance surface area
(p(r
o
2
r
i
2
)) in LCS and PLIF is calculated to be 3.22 and
3.75 cm
2
, respectively. The percent difference is compa-
rable to the repeatability of Reynolds number and L/D.
Both Shariff et al. (2006) and Shadden et al. (2006)
used attracting and repelling LCSs for an already formed
vortex ring to investigate the entrainment/detrainment
between irrotational uid outside the vortex ring and the
uid in the vortex ring. Here, however, we are interested in
the ow of uid into the developing spiral. While the
attracting LCS identies the spiral (i.e., unstable manifold
which separates ejected uid from entrained uid in the
vortex ring), the repelling LCS does not. Thus, we only use
the attracting LCS in this analysis. With the spiral entrance
identied, Q
E
is computed as
t / t
P
W
r
/
U
M
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
Fig. 10 Vortex ring velocity obtained from polynomial t of vortex
peak locations (Re
J
= 1,000, L/D = 2.0)
Exp Fluids (2008) 44:235247 243
1 3
Q
E
t 2p
Z
r
o
r
i
ur W
r
rdr 9
where u(r) is obtained from DPIV and W
r
is the ring
velocity. Although Dr considered here is different from the
spiral entrance identied in the PLIF results (AB in Fig. 8),
that should not affect the comparison of Q
E
between the
methods since the uid entering AB also passes through
Dr.
The comparison of Q
E
obtained from LCS/DPIV and
PLIF is given in Fig. 12. The average slope obtained from
PLIF data for t* 0.5 and t* 1.0 demonstrates reason-
able agreement with LCS/DPIV results for the same
interval. However, PLIF results during 0.5 \t* \1.0 were
not able to capture gradual increase of entrainment
obtained from LCS/DPIV. The disagreement is a combi-
nation of the fact that the PLIF result is an average value
and uncertainty in V
E
obtained by PLIF due to dye diffu-
sion in the vortex core. Nevertheless, the simplistic
approach used in PLIF results during ring formation give
good quantitative measurements of entrainment rates dur-
ing and after jet ejection, even though the transition
between the two rates appears more abrupt in the PLIF data
than is actually indicated by the LCS/DPIV data.
The high entrainment rate after the jet stops agrees with
the previous observation that most of the ambient uid is
entrained during the impulse preserving phase of motion.
As indicated in Fig. 13, a key factor in the increased Q
E
is
the dramatic increase in Dr after jet termination. From
Eq. (9), however, the ow velocity plays a role as well. In
particular, note that time-accurate LCS results show a
descending trend for 0.9 t* 1.5. The peak near t* = 0.9
is reasonable since piston velocity program starts to slow
down after this point. This causes reduction in velocity, but
the Dr is almost constant between t* of 1 and 1.5. Con-
sequently, the entrainment rate starts to diminish until Dr
starts to rise after t* of 1.5. This is observed as the LCS
entrainment rate starts to increase again at late time (i.e.,
t* [1.5).
Fig. 11 Entrance surface area
through which ambient uid is
entrained is shown for both LCS
data in a with |T| = 5.01 s
(|T|/t
p
= 1.81) and PLIF data in
b with t* = 1.81. Reference
horizontal lines based on the
LCS data are given for
comparison purposes
t / t
p
Q
E
/
(
0
.
2
5
p
D
P 2
U
M
)
Fig. 12 Entrainment rate comparison between PLIF and LCS/DPIV
results (Re
J
= 1,000 and L/D = 2.0)
244 Exp Fluids (2008) 44:235247
1 3
3.4 Quantitative comparison between trapezoidal
and triangular velocity programs
The quantitative validity of the PLIF results having been
conrmed, only PLIF entrainment results will be used here
for brevity. The effect of velocity program on entrainment
rate obtained from PLIF data can be seen in Fig. 14. Since
trapezoidal and triangular NS velocity programs exhibit
similar acceleration behavior at the start up, nearly the
same entrainment rates are obtained during jet initiation
(0 t* 1.0). Similarly, during the momentum conserving
period (1.0 t* 1.5), near identical entrainment rates are
calculated for trapezoidal and triangle PS velocity
programs since these programs behave similarly at the
deceleration phase.
For rapidly accelerating velocity programs, both the
vorticity in the forming spiral and the entrainment area are
key parameters. First, Fig. 15 shows that the fast acceler-
ation cases (i.e., trapezoidal and triangular NS) have higher
circulation during jet ejection, which produces a higher
entrainment velocity (due to the Biot-Savart induction)
compared to the triangular PS velocity program. Second,
since rapidly accelerated velocity programs cause tighter
spirals for 0 t* 1.0, a larger Dr is available for
entrainment as shown in Fig. 16.
For rapidly decelerating velocity programs, on the other
hand, the primary effect is on the area through which uid
is entrained (Dr), as determined by the effect of the stop-
ping vortex. Figure 16 shows that Dr rapidly increases for
both the trapezoidal and triangular PS cases following jet
termination, arriving at a nal value of nearly twice that of
the triangular NS case. This is contrasted with the total
circulation, C
M
, which differs by only about 30% between
these three cases.
Finally, to understand the effect of L/D on total
entrained uid volume, the entrainment fraction
i.e., g
ent

V
E
V
B
is plotted in Fig. 17 below. When Re
J
is
xed, the piston is required to reach the commanded
velocity in a shorter time as L/D is decreased. This results
in a higher g
ent
as L/D is decreased for all the velocity
programs given in Fig. 17 as a more compact vortex core is
generated at initiation for higher initial jet acceleration. On
t / t
P
D
r
/
D
0.25 0.5 0.75 1 1.25 1.5 1.75 2
0
0.05
0.1
0.15
0.2
0.25
Fig. 13 Variation of
Dr
D
for Re
J
= 1,000 and L/D = 2.0 (The
uncertainty of
Dr
D
is less than 5%)
t / t
P
Q
E
/
(
0
.
2
5
p
D
P 2
U
M
)
0.25 0.5 0.75 1 1.25 1.5 1.75 2
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0.2
0.22
Trapezoidal
Triangular NS
Triangular PS
Fig. 14 PLIF entrainment rate (calculated from the average slope of
PLIF entrained volume data) for trapezoidal, triangular NS and PS
velocity programs at Re
J
= 1,000 and L/D = 2.0
t / t
p

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6


0
0.2
0.4
0.6
0.8
1
1.2
Trapezoidal
Triangle (NS)
Triangle (PS)
Fig. 15 Vortex ring circulation comparison for trapezoidal, triangu-
lar NS, and PS velocity programs for Re
J
= 1,000 and L/D = 2.0.
Circulation is calculated from C
R
1
0
R
1
0
x
h
dxdr and C
M
refers the
maximum circulation obtained from trapezoidal velocity program.
Uncertainty of the circulation calculations is less than 1% of the
maximum circulation
Exp Fluids (2008) 44:235247 245
1 3
the other hand, the proximity of the ring to the nozzle when
the jet terminates and the strength of the stopping vortex
determine the nal size of Dr (Dr
F
) following jet termi-
nation. When L/D is decreased causing a stronger stopping
vortex, Dr
F
/D increases as illustrated in Fig. 18. This also
contributes to the larger g
ent
at small L/D.
4 Conclusions
Ambient uid entrainment during vortex ring formation
due to Biot-Savart induction was investigated from a
pistoncylinder mechanism. PLIF and DPIV combined
with LCS methods were utilized rst to identify the vortex
bubble and then to compute the ambient uid entrainment.
The piston velocity programs and L/D ratio were changed
in order to study the effects of these parameters on
entrainment in the forming vortex ring.
PLIF method gives entrainment during ring formation
using the assumption of the fore-aft symmetry. This
assumption is justied for a formed ring. During ring for-
mation it is also justied for calculation of V
E
due to
approximate fore-aft symmetry of the vortex spiral. The
DPIV streamline method accurately provides the vortex
bubble shape via the stagnation streamline once the vortex
is formed. Indeed, the streamfunction shows bubble growth
by diffusion; however, it can only be used after the ring is
formed. The DPIV/LCS method provides more detailed
information than either method, but requires signicantly
more data processing. The LCS results conrmed the
conclusions drawn from PLIF using the assumption of fore-
art symmetry.
Studying the PLIF results in detail revealed several key
factors affecting entrainment during vortex ring formation.
First, the effect of the velocity program on entrainment rate
is determined primarily by the magnitude of jet accelera-
tion during initiation and termination as shown in Fig. 14.
While high initial accelerations such as a trapezoidal or a
triangular NS velocity program can enhance the initial
entrainment rate by as much as 50% compared to a trian-
gular PS velocity program during jet ejection, high
terminal deceleration like a trapezoidal velocity program or
a triangular PS velocity program can increase the nal
entrainment rate up to only 10% compared to a triangular
t / t
P
D
r
/
D
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
0
0.025
0.05
0.075
0.1
0.125
0.15
0.175
0.2
0.225
Triangular NS
Triangular PS
Fig. 16 Dr comparison between triangular NS and PS velocity
programs for Re
J
= 1,000 and L/D = 2.0. The dashed line shows the
Dr
D
obtained from trapezoidal velocity program as a reference
L / D

e
n
t
0 0.5 1 1.5 2 2.5
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
Trapezoidal
Triangle (NS)
Triangle (PS)
Fig. 17 Entrainment fraction i.e.; g
ent

VE
VB
comparison for vortex
rings at Re
J
= 1,000
L / D
D
r
F
/
D
0 0.5 1 1.5 2 2.5
0.08
0.1
0.12
0.14
0.16
0.18
Fig. 18 Dr
F
/D variation with respect to L/D for Re
J
= 1,000 and the
Triangular NS velocity program
246 Exp Fluids (2008) 44:235247
1 3
NS velocity program. Second, L/D has a strong effect on
entrainment as well. As L/D is lowered from 2.0 to 0.5, the
piston must be accelerated at a higher rate causing higher
vorticity in the forming spiral. Also, when L/D is reduced
from 2.0 to 0.5, Dr
F
/D increases by as much as 80%. This
is due to the rapid stop of the piston which generates a
stronger stopping vortex and creates a larger area for
ambient uid entrainment. The net effect of these trends is
an increase of up to 67% in the nal entrainment fraction as
L/D is reduced from 2.0 to 0.5.
These observations provide insight into enhancing
ambient uid entrainment during vortex ring formation.
Specically, a trapezoidal velocity program with low L/D
ratio is an ideal candidate since this program benets from
an impulsive jet initiation as well as a rapid jet termination.
Acknowledgments This material is based upon work supported by
the National Science Foundation under Grant No. 0347958.
References
Akhmetov DG, Lugovtsov BA, Tarasov VF (1980) Extinguishing gas
and oil well res by means of vortex rings. Combus Explos
Shock Wave 16:490494
Auerbach D (1991) Stirring properties of vortex rings. Phys Fluids
A3:13511355
Dabiri JO, Gharib M (2004) Fluid entrainment by isolated vortex
rings. J Fluid Mech 511:311331
Didden N (1979) On the formation of vortex rings: rolling-up and
production of circulation. Z Angew Math Phys 30:101116
Fabris D, Liepmann D (1997) Vortex ring structure at late stages of
formation. Phys Fluids 9:28012803
Green S (1995) Fluid vortices. Kluwer, Dordrecht
Johari H (1995) Chemically reactive turbulent vortex rings. Phys
Fluids 7:24202427
Kercher DS, Lee JB, Brand O, Allen MG, Glezer A (2003) Microjet
cooling devices for thermal management of electronics. IEEE
Trans Components Packaging Technol 26:359366
Maxworthy T (1972) The structure and stability of vortex rings.
J Fluid Mech 51(part 1):1532
Maxworthy T (1977) Some experimental studies of vortex rings.
J Fluid Mech 81(part 3):465495
Shadden SC, Lekien F, Marsden JE (2005) Denition and properties
of Lagrangian coherent structures from nite-time Lyapunov
exponents in two-dimensional aperiodic ows. Physica D
212:271304
Shadden SC, Dabiri JO, Marsden JE (2006) Lagrangian analysis of
uid transport in empirical vortex ring ows. Phys Fluids
18:047105047111
Shariff K, Leonard A (1992) Vortex rings. Annu Rev Fluid Mech
24:235279
Shariff K, Leonard A, Ferziger JH (2006) Dynamical systems analysis
of uid transport in time-periodic vortex ring ows. Phys Fluids
18:047104047111
Westerwheel J, Dabiri D, Gharib M (1997) The effect of a discrete
window offset on the accuracy of cross-correlation analysis of
digital PIV recordings. Exp Fluids 23:2028
Willert CE, Gharib M (1991) Digital particle image velocimetry. Exp
Fluids 10:181193
Exp Fluids (2008) 44:235247 247
1 3

S-ar putea să vă placă și