Sie sind auf Seite 1von 4

PHYSICS OF FLUIDS 20, 041703 2008

Open-loop control of fully nonlinear self-excited oscillations


M. P. Hallberg and P. J. Strykowski
Department of Mechanical Engineering, University of Minnesota, Minneapolis, Minnesota 55455, USA

Received 25 January 2008; accepted 21 March 2008; published online 24 April 2008
Experiments conducted using a low-density axisymmetric jet reveal that the self-excited nature can
be altered by modulating the external pressure field using an acoustic driver. A region exists in the
forcing frequency-amplitude space where the self-sustaining frequency is entirely absent and solely
replaced by the forcing frequency and harmonics thereof. Inside this region, the centerline
streamwise velocity disturbance level can be significantly increased or decreased. Visual evidence
suggests these alterations coincide with dramatic changes in jet spreading and demonstrates that
forcing can provide an effective means of controlling mixing in self-excited jets. Results support
theoretical predictions made by Pier B. Pier, Proc. R. Soc. London, Ser. A 459, 1105 2003.
2008 American Institute of Physics. DOI: 10.1063/1.2908008
Compelled by recent theoretical insights suggesting the
roleor more accurately, lack thereofof external perturbations applied to self-excited flows may be overturned,1,2 a
study was undertaken in hopes of showing that self-excited
structures can be altered by external excitation. Self-excited
jets have long been shown to be insensitive to external
excitationa fact that is often used as evidence that a particular jet is indeed self-excited.35 This is in contrast to convectively unstable jets that are intimately connected to background and/or imposed forcing.6 These two distinct modes of
operation have led to the classification of systems as oscillators or amplifiers. Results presented herein suggest that the
former at least in the case of low-density jets is, in fact,
receptive to external excitation via harmonic acoustic pressure waves.
Open fluid systems are known to support self-sustaining
oscillations, and within these open systems, several key elements appear to be universal. First, there exists a particular
location where a linear wavemaker resides7the transition
from convective to absolute instability. The wavemaker is
preceded by a linearly decaying tail8 and followed by a nonlinear wavetrain tuned to the linear wavemaker frequency.7,9
As the name suggests, these dynamics are controlled by intrinsic features present in the system and have long been
believed to be independent of external influences. Recent
theoretical predictions by Pier1 using a model based on the
complex GinzburgLandau equation suggest that the inherent dynamics can be replaced if the proper forcing is applied.
Depending on the location of the applied forcing, the
strength necessary to suppress the self-sustaining process
may be exceedingly small. This technique has been analytically applied to the rotating boundary layer, where it was
shown that the onset to secondary absolute instability and
possibly transition to turbulence was delayed approximately
100 boundary layer units.2 The results that follow aim to
support these predictions by using experiments conducted in
a free jet flow capable of producing self-sustaining dynamics. Recent evidence in low-density10,11 and swirling jets12
supports the theoretical description above, further suggesting
that the open-loop control technique might indeed prove use1070-6631/2008/204/041703/4/$23.00

ful. Finally, early work on the suppression of vortex shedding by cylinder vibrations13 could ultimately be connected
to the same theoretical mechanisms.1
The experiments were conducted using a low-density jet
subjected to external excitation by using an acoustic driver to
add small perturbations to the flow. The jet consisted of a
fifth-order polynomial nozzle with a diameter contraction ratio of 10:1 D = 0.635 cm. After the contraction, a short extension three diameters was added to allow additional control over the boundary layer thickness at the jet exit. The end
of the extension piece was carefully machined to produce a
fine knife edge with a thickness much less than the exit diameter to allow free entrainment. The gas used to drive the
flow was helium, it exhausted into ambient air producing a
density ratio of 0.14. This was chosen because low-density
jets have been shown to support self-sustaining oscillations
of the type accurately modeled by the model equation.4,14
The flow was controlled by fixing the mass flow of helium,
thereby setting the Reynolds number Re= U0D / based on
jet centerline conditions which for most results presented
herein was set at 1130, corresponding to D / 0 = 30.9, where
0 is the momentum thickness at the jet exit integrated to
ur / U0 = 0.2. Briefly see Refs. 35, 10, 11, 15, and 16 for
detailed observations, above the onset to self-sustained oscillations, the low-density jet is marked by a rather short
potential core followed by a rapid and at times dramatic radial expulsions of fluid breakdown to turbulence as can be
seen in the schlieren image of Fig. 1a. A Dantec 55P11
hotwire located two diameters downstream and positioned on
the centerline was used to produce the spectrum in Fig. 2a
which illustrates the pure-tone oscillation present in the
flowa feature long associated with the self-excited nature
of the flow, is independent of position i.e., global, and for
the present jet is 818 Hz St = f gD / U0 = 0.234, where f g is the
peak in the unforced spectrum and is hereafter the global
frequency.
Next, a 13.3 cm speaker was located 30.5 cm from the
jet exit at an angle of 40 with respect to the jet centerline.
The speaker cone was directly aimed at the jet exit such that
for a given spatial separation, the pressure fluctuations at the

20, 041703-1

2008 American Institute of Physics

Downloaded 02 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

041703-2

Phys. Fluids 20, 041703 2008

M. P. Hallberg and P. J. Strykowski

(c)

(a)
0

1000

2000
3000
Frequency (Hz)

4000

5000

1000

2000
3000
Frequency (Hz)

4000

5000

(b)
0

FIG. 2. Power spectrum of a unforced globally unstable helium jet a and


spectrum of jet exposed to external acoustic forcing at 1227 Hz f f / f g
= 1.5 and 109 dB b.
112
1.3

1.1

1.

0.2

0.6

108
2

0.4
0.8

1.1

104
0.4

100

0.6
1

jet exit would be maximized. The speaker was driven by a


stereo amplifier and fed a sinusoidal signal from signal generating software. The frequency and amplitude of the signal
were readily varied, and the resulting response of the flow
was monitored by using a hotwire located two diameters
downstream on the jet centerline. A brief sweep of forcing
frequency f f and amplitude space produced a sea of combinations capable of altering the jet dynamics. One particular
caseby no means uniqueis illustrated. The speaker was
driven at a frequency of 1227 Hz and amplitude of 109 dB
sound pressure level. Sound pressure levels were measured
at the nozzle exit by using an American Recorder SPL-8810
sound pressure level meter with C-weighting. The resulting
flow can be seen in Fig. 1b along with the power spectrum
in Fig. 2b. Obvious changes can be seen in both Figs. 1 and
2 by the addition of the external acoustic forcing. Namely,
the schlieren imaging indicates a reduction in the jet mixing
i.e., longer potential core and shallower spreading, while
the hotwire spectrum reveals that the self-excited frequency
has been completely replaced by the forcing frequency while
simultaneously reducing the energy in the spectrum. While
previous authors have looked at the role of forcing on selfexcited dynamics in low-density jets,35 no one has reported
reductions in disturbance levels, rather they have hinted only
at increases.3,5
The last observation the overall lowering of the fluctuating amplitude, while it may be intriguing, is merely fortuitous since the choice of forcing yielded a reduction in amplitude and hence a reduction in mixing. Upon closer
investigation, a range of control was found to be possible, in
some cases the mixing was reduced as in Fig. 1b, in others
it was enhanced see, e.g., Fig. 1c. Varying degrees of
control
were
found
to
occur
from
roughly
0.25 f f / f g 1.9though the exact range is likely outside
that achieved with the current setup, which was limited by
the voltage at which distortion occurred 10 Vrms. A contour
map was created to illustrate the range of control achieved
and can be seen in Fig. 3. Contours of constant forced centerline fluctuation intensity urms,f measured at two diameters normalized by the unforced centerline globally unstable
intensity urms,g, also measured at two diameters, point to
the large differences achievable by using the simple control
authority. While the measurement location chosen is somewhat arbitrary, Fig. 4 reveals that the unforced disturbance
reaches its maximum amplitude near two diameters. Additionally, if considering jet mixing control as in Fig. 1 for a
given geometry, fixing the probe position would seem most
apropos.

Log Power

FIG. 1. Schlieren flow visualization of a globally unstable helium jet without a and with acoustic forcing at b 1227 Hz f f / f g = 1.5, and 109 dB
and c 730 Hz f f / f g = 0.89 and 115 dB.

Log Power

(b)

SPL (dB)

(a)

0.8

96

92

0.4

0.6

0.8

1.2

1.4

1.6

1.8

ff / fg

FIG. 3. Control map showing contours of constant forced streamwise fluctuations measured at two diameters urms,f normalized by unforced global
fluctuations at 2D urms,g. Inside the dashed contour, the global frequency is
absent and replaced by the forcing frequency as in Fig. 2b. Shaded region
exceeds 10 Vrms system limit.

Downloaded 02 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

041703-3

Phys. Fluids 20, 041703 2008

Open-loop control

urms/Uo

10

10

0.12
0.1

0.06

rms

/U

0.08

0.04
0.02
0

10

0.5

1.5

X/D

FIG. 4. Spatial evolution of the streamwise velocity fluctuations along centerline of unforced and forced jet f f / f g = 1.5 and 109 dB.

Superimposed on the contour map of Fig. 3 is a dashed


line that traces out the region where the peak in the power
spectrum is no longer dominated by the global frequency,
and generally has been replaced by the forcing frequency as
in Fig. 2b. When compared to the spectrum reported by
Sreenivasan et al.3 showing evidence of lock-in, a fundamental difference can be seen. The lock-in spectrum shows
many integer rational multiples of the excitation frequency
which includes the global frequency. The current results indicate that once inside the dashed contour only evidence of
the excitation frequency and its harmonics exist. Outside this
contour, there is still a significant amount of energy contained within the global frequency, resulting in complex nonlinear interactions between the inherent and the imposed frequencies producing power spectrum more reminiscent of the
lock-in spectrum of Ref. 3.
It should be noted that the details presented ultimately
depend on several factors. The choice of operating condition
was somewhat arbitrary but selected to provide a balance
between the degree of supercriticality5,15 of the jet and the
useful limits of the acoustic system. However, these results
are not unique to this operating condition. It was observed
that at a significantly higher Reynolds 1500 corresponding to a global frequency of 1200 Hz, a large reduction in
fluctuation amplitude was observed when the jet was forced
at 1500 Hz and 93 dB. Further, the precise method used to
generate the jetwhether using a pure gas as in the present
results or through heating air16will likely alter the precise
details. Last, the off-axis acoustic setup will produce frequency dependent phase differences across the nozzle up to
11% for the present study; however, as a check, the speaker
was relocated to provide axisymmetric forcing for the case
shown in Fig. 1b. Schlieren imaging revealed similar results at both speaker locations, suggesting that the off-axis
setup is of secondary importance.
The spatial evolution of the centerline velocity fluctuation was mapped out for the flows in Fig. 2. A Dantec 55P11
hotwire was positioned along the centerline at the jet exit and
axially traversed. The fluctuating velocity was determined

from 10 s traces sampled at 10 kHz. The hotwire was calibrated in helium and, provided the core remained intact,
yielded unambiguous velocity information.5 Figure 4 reveals
the spatial evolution of the centerline streamwise velocity
fluctuation intensity for the low-density jet when the flow is
undergoing natural self-sustaining oscillations as well as
when the flow is forced above the critical level outlined in
Fig. 3. Interestingly, throughout the jet the addition of the
forcing results in a lowering of the fluctuation intensity along
the centerline. At first glance, this result might seem paradoxical. Namely, why should a flow considered to be an
oscillator yield so dramatically to an external influence?
There exists a theoretical explanation for the preceding paradox derived by Pier1 using the complex GinzburgLandau
equation. Pier showed that through the addition of a forcing
term, a self-excited flow can be switched off with the proper
choice of forcing. Subsequently, the flow is free to evolve
based on the new user-supplied forcing, one that may or may
not yield improvements i.e., control from a fluid mechanical standpoint. This prediction is entirely consistent with Fig.
3 in that a domain exists where the flow is governed by the
user-selected open-loop forcing, and within that region the
centerline fluctuation intensity may be increased or decreased depending on the exact choice of forcing conditions.
Further predictions from the theory of Pier1 can also be
checked in the present study. First, the theory is based largely
on the location where nonlinearities start to augment the
flow. In order for the flow to be controlled by the imposed
forcing, the location where nonlinearity begins must be at or
upstream of the natural location. The near-field evolution
shown in Figs. 1 and 4 reveals that the forced flow is immediately yielding upon the jet exit, whereas the self-excited jet
displays nonlinear characteristics at approximately x = D / 2.
Second, the theory posits that the imposed spatial growth
should be less than the unforced natural growth; once again,
Fig. 4 reveals that the forced growth in the near-region x
D / 2 of the exit is lower than that of the self-sustained
spatial growth. Lastly, theory suggests that forcing levels required to achieve such an effect can be made exponentially
small provided the location of the imposed disturbance can
be far enough upstream preceding the convective-absolute
transition. Unfortunately, the present setup does not allow
one to probe the last conjecture. Incoming acoustic waves are
immediately converted to vortical waves at the jet exit,17
hence without decoupling the disturbance location and the
convective to absolute transition, some uncertainty regarding
forcing levels remains. Please note that, provided these connections to theory are real, localization of a spatially growing
harmonic disturbance upstream of the jet exit could produce
the same result with far less energy input.
In conclusion, experimental investigations into the role
of external acoustic forcing on the intrinsic dynamics of a
low-density jet reveal that the flow can be dramatically
alteredincluding both a significant increase in velocity disturbance intensity as well as an even more dramatic decrease
in intensity associated visually with a near relaminarization
of the jet. The result is a function of forcing frequency and
amplitude and exists over a significant range of conditions

Downloaded 02 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

041703-4

M. P. Hallberg and P. J. Strykowski

believed to be limited in the present study by the limits of the


acoustic setup. Due to the rather simple nature of the control,
it is conceivable that the approach could provide a useful
tool to domains that are susceptible to global instabilities.
While somewhat tenuous connections to the spatial growth
predictions of Pier1 exist, the absence of the self-excited frequency and rational multiples thereof and concomitant emergence of the forcing frequency are well aligned with theoretical predictions and demonstrate as envisioned by Pier that
the application of pure-tone forcing could provide a useful
control strategy. Ultimately, localization of the harmonic disturbance upstream of the jet exit could prove exceedingly
powerful.
The authors would like to acknowledge support from the
National Science Foundation under Grant No. CTS-0317429
as well as University of Minnesota support provided through
a doctoral dissertation fellowship.
1

B. Pier, Open-loop control of absolutely unstable domains, Proc. R. Soc.


London, Ser. A 459, 1105 2003.
2
B. Pier, Primary crossflow vortices, secondary absolute instabilities and
their control in the rotating-disk boundary layer, J. Eng. Math. 57, 237
2007.
3
K. R. Sreenivasan, S. Raghu, and D. Kyle, Absolute instability in variable density round jets, Exp. Fluids 7, 309 1989.
4
P. A. Monkewitz, D. W. Bechert, B. Barsikow, and B. Lehmann, Self-

Phys. Fluids 20, 041703 2008


excited oscillations and mixing in a heated round jet, J. Fluid Mech. 213,
611 1990.
5
D. M. Kyle and K. R. Sreenivasan, The instability and breakdown of a
round variable-density jet, J. Fluid Mech. 249, 619 1993.
6
S. C. Crow and F. H. Champagne, Orderly structure in jet turbulence, J.
Fluid Mech. 48, 547 1971.
7
B. Pier and P. Huerre, Nonlinear self-sustained structures and fronts in
spatially developing wake flows, J. Fluid Mech. 435, 145 2001.
8
B. Pier, P. Huerre, J.-M. Chomaz, and A. Couairon, Steep nonlinear
global modes in spatially developing media, Phys. Fluids 10, 2433
1998.
9
J.-M. Chomaz, Global instabilities in spatially developing flows: Nonnormality and nonlinearity, Annu. Rev. Fluid Mech. 37, 357 2005.
10
L. Lesshafft, P. Huerre, P. Sagaut, and M. Terracol, Nonlinear global
modes in hot jets, J. Fluid Mech. 554, 393 2006.
11
J. Nichols, P. Schmid, and J. Riley, Self-sustained oscillations in variabledensity round jets, J. Fluid Mech. 582, 341 2007.
12
F. Gallaire, M. Ruith, E. Meiburg, J.-M. Chomaz, and P. Huerre, Spiral
vortex breakdown as a global mode, J. Fluid Mech. 549, 71 2006.
13
M. Schumm, E. Berger, and P. A. Monkewitz, Self-excited oscillations in
the wake of two-dimensional bluff bodies and their control, J. Fluid
Mech. 271, 17 1994.
14
S. Raghu and P. A. Monkewitz, The bifurcation of a hot round jet to
limit-cycle oscillations, Phys. Fluids A 3, 501 1991.
15
M. P. Hallberg and P. J. Strykowski, On the universality of global modes
in low-density axisymmetric jets, J. Fluid Mech. 569, 493 2006.
16
L. Raynal, J.-L. Harion, M. Favre-Marinet, and G. Binder, The oscillatory instability of plane variable-density jets, Phys. Fluids 8, 993 1996.
17
D. Bechert and E. Pfizenmaier, Optical compensation measurements on
the unsteady exit condition at a nozzle discharge edge, J. Fluid Mech. 71,
123 1975.

Downloaded 02 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Das könnte Ihnen auch gefallen