CFD Simulations of Acoustic and Thermoacoustic Phenomena in Internal Flows OPEN FOAM
CFD Simulations of Acoustic and Thermoacoustic Phenomena in Internal Flows OPEN FOAM
AIAA Aviation
13-17 June 2016, Washington, D.C.
46th AIAA Fluid Dynamics Conference
Esteban D. Gonzalez-Juez∗
Combustion Science & Engineering, Inc., Columbia, MD.
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
I. Introduction
Acoustic and thermoacoustic phenomena in internal subsonic flows arise in many engineering applications.
An everyday example is the destructive interference of acoustic waves occurring in automotive mufflers.1 An-
other one is the interaction of turbulent shear layers with acoustic waves in flows past cavities.2 Examples of
these cavities include airplane wheel wells, nooks for measuring devices, and the entire passenger compart-
ment of a car when the side windows are open.3 Acoustic waves are also produced by temperature variations.
This happens, for example, from sudden temperature changes produced during filling in cryogenic applica-
tions; due to the temperature fluctuations associated with the flame in many combustion systems, such as
gas turbine engines and furnaces; and from temperature gradients in heat exchangers of thermoacoustic
engines and refrigerators. If the physical mechanism generating the acoustic waves in these applications
resonate with the acoustics of the corresponding enclosure, very large pressure fluctuations can be produced,
something desirable in a thermoacoustic engine, but a serious hazard to a gas turbine engine. It is this inter-
action of acoustic waves with an enclosure what distinguishes the occurrence of acoustic and thermoacoustic
phenomena in an internal flow versus an external one. In all of these applications, and others as well, it is
desirable to be able to predict these phenomena, either for the purpose of maximizing the acoustic noise from
a device, as in a thermoacoustic engine, or minimizing it, in order to reduce noise and structural vibrations.
With this motivation in mind, the present work evaluates the capability to capture acoustic and thermoa-
coustic phenomena in internal subsonic flows of two computational-fluid-dynamic (CFD) solvers from the
OpenFOAM
R library.4 OpenFOAM is an increasingly popular open source C++ library for the solution of
partial differential equations using the finite volume technique. It contains a vast number of physical models
for the simulation of flows, including thermodynamic and material property databases. Thus it is more than
just a CFD code. Its latest official version, 2.1.1 at the time of this writing, includes various CFD solvers.
Two of these are tested here: rhoPimpleFoam and rhoCentralFoam.
The expectation for this work is to belong to a series of ongoing studies that are evaluating Open-
FOAM and constitute, as a whole, a method of quality control of the code. So far, these studies have con-
ducted evaluations in a variety of flow problems: bluff-body flows;5–7 high-speed steady flows;8 multiphase
flows;9 cavitation;10 gas dispersion;11 diesel-engine combustion;12 combustion in aerospace applications;13
and flames.14–16 However, evaluations of OpenFOAM for acoustic applications are scant17, 18 and, to the
author’s knowledge, none of these are as comprehensive as the present work.
∗ Senior Engineer, [email protected], AIAA Member.
Copyright © 2016 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
1 of 21
The current study considers four different types of acoustic phenomena: the propagation of acoustic
waves with no mean flow (Sec. III.A); the production and propagation of acoustic waves due to impulsive
heating (Sec. III.B); hydrodynamic/acoustic interactions in a cavity flow (Sec. III.C); and, thermoacoustic
instabilities (Sec. III.D). Each of these phenomena is simulated by using configurations selected for their
practical relevance and simplicity, so that others could reproduce the present results with a small computer
cluster. In what follows, predictions with rhoPimpleFoam and rhoCentralFoam are compared with data from
experiments and/or other CFD codes.
single-component:
∂ρ ∂(ρui )
+ =0, (1)
∂t ∂xi
∂ρui ∂(ρuj ui ) ∂p ∂
+ =− + (τji + Tji ) , (2)
∂t ∂xj ∂xi ∂xj
∂ρH ∂(ρuj H) ∂ ∂p
+ =− (qj + Qj ) + , (3)
∂t ∂xj ∂xj ∂t
2 of 21
rhoPimpleFoam is a transient, implicit, pressure-based solver. As indicated in Tab. 1, rhoPimpleFoam
uses an inner iterative loop to correct the velocity field using the output of a pressure equation, as in
the well-known PISO algorithm,21 and also uses an outer iteration loop for additional corrections. A good
discussion of the implementation of the PISO algorithm in OpenFOAM is given in Jasak.20 Various temporal
and spatial discretization options are available for rhoPimpleFoam. In the following, unless said otherwise,
gradients and Laplacians are discretized with central schemes, and convective fluxes are discretized using a
TVD scheme with van Leer flux limiter. With this limiter, the flux limiter function φ is given by:22
r + |r|
φ= , (6)
1 + |r|
with r being the ratio of successive gradients on the mesh for a particular field. Another limiter function
used in the present work is OpenFOAM’s limited linear function,
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
2r
φ = max 0, min 1, , (7)
k
where k is a constant. Time is discretized with a second-order backward-differencing scheme that uses the
current and previous two time-step values.
Table 1. Outline of the algorithm used by rhoPimpleFoam.
Calculate density
Start outer loop
Compute uncorrected velocity with momentum eqn.
Compute enthalpy with energy eqn.
Start inner loop
Get density from eqn. of state
Get pressure with pressure eqn.
Get density from eqn. of state
Correct velocity
End inner loop
Calculate turbulent quantities
End outer loop
rhoCentralFoam is a transient, density-based solver that uses a time-splitting method to solve the above
equations in two steps: First, it advances explicitly the governing equations with inviscid terms only in a
predictor step, and then it does so implicitly with the diffusion terms in a corrector step. Another distin-
guishing feature of rhoCentralFoam is that it uses, instead of approximate Riemann solvers, the Kurganov-
Noelle-Petrova23, 24 central-upwind scheme. An in-depth description of rhoCentralFoam and a corresponding
thorough validation study for high-speed flows is provided in Greenshields et al.8 Unless otherwise stated,
rhoCentralFoam is used with a second-order time-discretization scheme.
With OpenFOAM, the solution of linear systems of algebraic equations can be achieved with various
methods, such as preconditioned conjugate and biconjugate gradient methods, and a generalized geometric-
algebraic multigrid method. The first two are used in the present work.
The formulation of the various boundary conditions used here (Neumann, Dirichlet and mixed conditions)
is straightforward, except for OpenFOAM’s (nonperfectly) nonreflective boundary condition, termed wave-
Transmissive. This boundary prevents exiting waves from reflecting back into the computational domain by
enforcing the following equation at a boundary:
∂p ∂p
+U = K(p∞ − p) , (8)
∂t ∂xn
where U = uc + c0 , uc is a characteristic convective speed, c0 is a reference speed of sound, xn is normal
to the boundary, K = U/l∞ , and p∞ is the pressure outside the boundary, evaluated at a distance l∞ . At
values of l∞ that are small enough, this boundary condition acts as a reflective boundary condition, whereas
for large enough values it is nonreflecting. More precisely, this boundary condition acts as a low-pass filter
3 of 21
with cutoff frequency of βK,25 with β being a constant of order 1. Care should be taken in selecting values of
p∞ and l∞ that do not produce drifts of the mean pressure inside the computational domain, cf. Sec. III.C.
This boundary condition is evaluated in the following.
III. Results
III.A. Acoustic waves in a muffler
The practical relevance and simplicity of the expansion-chamber muffler make it a well-suited problem to test
the ability of a CFD solver to capture acoustic phenomena. The 2D-axisymmetric geometry shown in Fig. 1 is
considered here. This geometry and some of the next parameters are taken from Middelberg et al.1 The tube
is long enough so that potential reflections at the inlet and outlet do not contaminate observables of interest
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
during the simulation. This allows one to focus on the solver performance rather than potential problems
with nonreflecting boundary conditions, which are dealt with later. The following boundary conditions are
used. At the inlet, the pressure gradient normal to it is zero, and the velocity undergoes an initial (single)
sinusoidal variation of amplitude 0.05 m/s and period 3.125e-4 s, which is shown later in Fig.2. After
this pulse, the velocity remains zero. Having a zero bulk velocity inside the muffler simplifies the analysis
considerably. At the outlet, zero-normal-gradient boundary conditions for velocity and pressure are used.
Walls are no-slip and adiabatic. Turbulence modeling is not needed. The working fluid is air at 101325 Pa
and 300 K. Meshes with uniform grid spacings of 4 mm and 2 mm are used and results for the latter are
discussed here; such spacings are satisfactory for the frequencies of interest here. The simulations are run
for 0.05 s with a time step of 2e-7 s (1e-7 s) for the 4 mm (2 mm) mesh. With rhoPimpleFoam, convective
fluxes are discretized with OpenFOAM’s limited linear scheme, cf. Eq. 7.
Figure 1. Schematic of the setup for the muffler simulations. Dimensions are in mm.
Expansion mufflers reduce the amplitude of high frequency components of the noise via a destructive
interference of sound waves in the expansion. Thus, a key observable is the transmission loss:26
P̂i
T L = 20 log10 , (9)
P̂t
where P̂i (P̂t ) is the Fourier transform of the pressure versus time signal measured at the centerline and
x=-7830 mm (x=294 mm). These pressure signals are shown in Fig. 2 for a representative case. Care must
be taken in selecting the exact same time interval to calculate T L for different simulations and experiments.
Figure 3 compares results from experiments, rhoCentralFoam, and rhoPimpleFoam. Notice that these solvers
capture the locations where T L = 0, corresponding to the natural frequencies of the expansion: 324, 648,
972, and 1296 Hz, for the first four modes. These frequencies can be estimated by assuming a 1/2-wave
mode inside the expansion, which gives f = nc/(2l), with n being an integer, c the sound speed (350 m/s)
and l the length of the expansion (540 mm). Figure 3 also shows that OpenFOAM solvers also capture
well the peak near 2700 Hz. This is important because solvers that are too dissipative predict a larger peak
(more losses) than the experiments. The largest disagreement between experiments and simulations occur
in the 0-200 Hz region, and may be caused by having a time signal that is not long enough to capture the
lower frequency phenomena. (The time traces obtained from the simulations for the postprocessing are 0.04
4 of 21
s long and, hence, include only 8 cycles of a 200 Hz component). In spite of this disagreement, the overall
agreement between experimental results and those obtained with rhoCentralFoam, and rhoPimpleFoam is
good. It should be kept in mind, however, that the grid spacing used (2 mm) would make the simulation of
actual mufflers very expensive. This issue highlights a well known challenge in computational aeroacoustics,
and points to the need of parallel clusters to simulate these problems.
5
x 10 (a) 5
x 10 (b)
1.0134 1.0133
1.0134
1.0133
1.0133
1.0133 1.0133
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
P (Pa)
P (Pa)
1.0132
1.0132 1.0132
1.0131
1.0132
1.0131
1.0131 1.0132
0 0.5 1 0.02 0.025 0.03 0.035 0.04
t (s) −3 t (s)
x 10
Figure 2. Pressure time traces Pi (t) (a) and Pt (t) (b) for the geometry of Fig. 1 obtained with rhoCentralFoam from
which the transmission loss T L is calculated.
To diagnose the boundary conditions, results for T L using the geometry of Fig. 1 are compared with
two other geometries. The first one is similar to that of Fig. 1 but truncated at the x=400 mm plane, at
which a nonreflective boundary condition, Eq. 8, is prescribed. The second geometry adds to this truncated
geometry a region of 424 mm radius between the x=400 mm and x=800 mm planes, and a nonreflective
boundary at x=800 mm. The meshing in this region is coarser in the streamwise direction. Such type
of region is usually called a sponge zone. A comparison between the T L vs. frequency curves obtained
with these two new setups with those from the original setup is given in Fig. 4. Only rhoCentralFoam
results are shown. With nonreflecting boundary conditions, predictions using the original domain should be
indistinguishable from those using the other domains. Figure 4 shows that using a nonreflective boundary
condition with the truncated domain degrades the T L vs. frequency curve. In particular, notice that the
peak at 2700 Hz gets damped. This degradation was observed at different values of the characteristic length
l∞ , cf. Eq. 8, in the range 1-20 m. Adding a sponge zone to the truncated domain, as can be seen in Fig. 4,
improves this degradation. From these results, two recommendations can be made to use OpenFOAM
for acoustic simulations that need nonreflective boundary conditions. First, OpenFOAM’s nonreflective
boundary condition should be used in combination with sponge zones. Second, since the design of sponge
zones can be cumbersome, it might be a good idea to design new boundary conditions, such as those based
on the NSCBC technique,27 as already done by Piscaglia et al.17
5 of 21
(a)
40
30
TL
20
10
0
0 500 1000 1500 2000 2500 3000
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
(b)
40
30
20
10
0
0 500 1000 1500 2000 2500 3000
f (Hz)
Figure 3. Transmission loss for the geometry of Fig. 1 predicted with rhoCentralFoam (a) and rhoPimpleFoam (b)
with a mesh spacing of 2 mm. Simulation results are shown with thin lines and experimental ones with thick lines.
(a)
40
30
TL
20
10
0
0 500 1000 1500 2000 2500 3000
(b)
40
30
20
10
0
0 500 1000 1500 2000 2500 3000
f (Hz)
Figure 4. Comparison of transmission losses predicted with rhoCentralFoam with the setup shown in Fig. 1 (solid thick
line) and a similar one but truncated at the x=400 mm plane and with an outlet boundary condition prescribed as
nonreflective (a) or by using a sponge outlet with nonreflective conditions (b).
6 of 21
heat capacity and temperature-dependent viscosity and thermal conductivity according to Sutherland’s law,
cf. Eq. 5. This last consideration is important since results with a constant viscosity did not agree as well
with those from Aktas.29 In contrast with the numerics used in the present work, cf. Sec. II, Aktas29 uses
explicit finite differencing and discretizes diffusion terms with central differences and the convection with
the flux-corrected transport algorithm.31
Figure 5 shows pressure contours obtained with rhoCentralFoam some time after the wave has been
reflected from right wall (t = 1.415e-5 s). Note that the flow is predominantly one-dimensional, and no
spurious oscillations are produced at the boundaries, a desirable result. The variation of pressure in the
horizontal direction at different times can be better seen in Fig. 6. As seen in previous experiments32 and
simulations,29 Fig. 6 shows that the pressure waves have a sharp front, result of the impulsive heating, with
a long tail following the front. These observations, discussed here for rhoCentralFoam, were also seen with
rhoPimpleFoam. A comparison of predictions from these solvers and from the highly-accurate code used by
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
Aktas29 is shown in Fig. 7, which shows plots of the pressure at the center of the square cavity versus time.
Notice in particular that as time increases pressure oscillations are attenuated, a result that is expected due
to diffusive processes. Figure 7 shows that results from rhoCentralFoam and rhoPimpleFoam agree fairly well
with those from Aktas,29 a disagreement being the prediction of larger peak-to-peak pressure oscillations by
OpenFOAM’s solvers. A difficulty in achieving a better agreement is that, since the present results are seen
to be very sensitive to the viscosity and heat conductivity, Aktas29 does not provide the exact way in which
these quantities were modeled in that work.
Figure 5. Instantaneous spatial distribution of pressure (in Pa) predicted with rhoCentralFoam after the wave gets
reflected from the right wall at t = 1.415e-5 s.
For the present thermoacoustic simulations, it was necessary to use time intervals an order of magnitude
smaller than those used by Aktas.29 By increasing the original time interval by a factor of five, rhoCentral-
Foam produces numerical instabilities, and rhoPimpleFoam reproduces previous results but with no savings
in computational cost, since it needed more iterations per time step to prevent numerical instabilities. These
results are not surprising since Aktas29 uses a more accurate code exclusively designed for the simulation
of shocks and acoustic waves in simple setups, whereas OpenFOAM CFD solvers are multi-purpose and
intended to be used in industrial applications. Nonetheless, there is some room for improving the computa-
tional efficiency of OpenFOAM CFD solvers, as discussed later on.
7 of 21
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
Figure 6. Horizontal variation of pressure taken at a horizontal plane intersecting the center of the square predicted
with rhoCentralFoam: t = 1.06126e-5 s (dashed line, wave moving from left to right), 1.2735e-5 s (dash-dot line),
1.415e-5 s (dash-dash-dot line, wave moving from right to left), and 1.48574e-5 s (solid line).
5
x 10 (a)
1.04
1.03
p (Pa)
1.02
1.01
0 0.5 1 1.5 2 2.5 3
t (s) x 10
−5
x 10
5 (b)
1.04
1.03
1.02
1.01
0 0.5 1 1.5 2 2.5 3
−5
x 10
Figure 7. Temporal variation of the pressure at the center of the square predicted by Aktas29 (thick line), rhoCentral-
Foam (a) and rhoPimpleFoam (b).
8 of 21
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
(no-model) LES. The fluid is air at 300 K and 1 atm. The top and bottom walls are no-slip and adiabatic.
Because of the high Reynolds number of this flow, and the lack of strong and adverse pressure gradients in
regions of interest, equilibrium wall functions are used. At the inlet, the velocity is spatially uniform and
equal to (260,0,0) m/s with white noise fluctuations (as done in2 ), the temperature is 300 K, and the gradient
of pressure normal to the inlet plane is zero. Currently there are better ways to prescribed inlet fluctuations
than using white noise, but this approach is the one used by Larcheveque et al.,2 and serves the present
goals. The Reynolds number based on the inlet velocity and the length of the cavity (L=50mm) is 8.6e5. At
the outlet, with rhoCentralFoam, an advective boundary condition is used for velocity and temperature, and
a nonreflecting boundary condition is used for pressure. Simulations with rhoCentralFoam using a sponge
zone at the outlet of the computational domain exhibit similar results to the ones discussed next. The
use of these outlet boundary conditions with rhoPimpleFoam produced a mean pressure inside the domain
that increased with time. This undesirable condition was corrected by using outlet boundary conditions for
velocity, temperature and pressure that vary, depending on the direction of the flow, between a zero gradient
normal to the outlet plane and a fixed value. (These are the so-called inletOutlet and outletInlet boundary
conditions in OpenFOAM.) Periodic (cyclic) boundary conditions are used in the transverse direction. The
mesh has a vertical grid spacing of 0.25 mm along the y=0 plane, that stretches in the horizontal direction to
2 mm (6 mm) at the inlet (outlet), and a horizontal grid spacing of 0.5 mm along the x=0 plane, stretching
to 3 mm at the top and bottom walls. The grid spacing in the transverse direction is constant and equal to 5
mm. This results in a total of 236000 cells; Larcheveque et al.2 uses meshes with 690110 and 1627320 cells.
Note that the final mesh used here is a result of various iterations. The simulations are run with a time step
of 1e-7 s. In comparison with the numerics used in the present work, cf. Sec. II, Larcheveque et al.2 uses a
finite-volume code, discretizes convective fluxes with either a centered scheme with corrective upwinding or a
MUSCL scheme of the AUSM+(P) family, and time advances the simulations with a third-order Runge-Kutta
scheme.
Vertical profiles of the mean horizontal and vertical velocities at the top of the cavity, both near the front
and the back, are shown in Fig. 9. Overall, both rhoCentralFoam and rhoPimpleFoam agree well with the
experiments.33 The largest disagreement is in the horizontal velocity profile in the region 0 < y/L < 0.05.
The cause of this discrepancy might be the use of a plug inlet profile for velocity rather than using a fit to
the experimental profile at the plane x = −L as done by Larcheveque et al.2 For the present purposes, which
9 of 21
is to determine if rhoCentralFoam and rhoPimpleFoam capture the mechanism of hydrodynamic/acoustic
interaction of interest here, the agreement in Fig. 9 is acceptable. Further support for the adequacy of the
turbulence modeling used is given in the Appendix.
(a) (b)
0.2
1
0
u/uinlet
v/uinlet
0.5
−0.2
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
−0.4
0
−0.2 0 0.2 −0.2 0 0.2
y/L
(c) (d)
0.2
1
0
u/uinlet
v/uinlet
0.5
−0.2
−0.4
0
−0.2 0 0.2 −0.2 0 0.2
Figure 9. Vertical profiles of the horizontal velocity u and vertical velocity v taken at the x=0.05L (squares) and
x=0.95L (circles) planes. Results from rhoCentralFoam (a,b), rhoPimpleFoam (c,d), and experiments33 (symbols) are
shown.
A predominant feature of this flow is the shear layer at the top of the cavity (x=0 plane). This feature
is highlighted in Fig. 10 with a vorticity contour. Notice the roll up of a vortex and impingement into the
back side of the cavity. As a result of this impingement, pressure (acoustic) waves are produced, as can be
seen in Fig. 10. These acoustic waves, in turn, affect the shear layer (that “produces” the vortices), closing
a feedback loop. This feedback mechanism is called the Rossiter mechanism and is captured by the present
simulations. The fundamental frequency of this mechanism is found to be 2179 in rhoCentralFoam, 2067
with rhoPimpleFoam, 2025 in the LES of Larcheveque et al.,2 and 1990 in the experiments.33 To provide
more details, Fig. 11 compares the pressure spectra obtained with present simulations with that from the
experiments. It can be seen in Fig. 11 that, for the first two peaks, the current simulations underpredict
the SPL amplitude by up to 8% and that, while rhoPimpleFoam predicts very well the frequencies, rhoCen-
tralFoam overpredicts them by up to 5%. With a larger time step (8e-7 instead of 1e-7 s), rhoPimpleFoam
needed more iterations per time step to produce convergent results, and produced a spectra as good as that
shown in Fig. 11. For the present type of phenomena, this agreement between OpenFOAM solvers and
experiments is very good, although not as good as that obtained by Larcheveque et al.,2 who used a finer
grid spacing. Table 2 compares the frequencies of various spectral peaks obtained with OpenFOAM’s solvers
and the code by Larcheveque et al.2
10 of 21
(a) (b)
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
(c) (d)
(e) (f)
Figure 10. Instantaneous spatial distribution of the pressure with a superimposed vorticity contour at intervals of 5e-5
s, obtained with rhoCentralFoam.
11 of 21
(a)
160
SPL (dB/Hz)
140
120
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
100
140
120
100
Figure 11. Pressure spectra measured at x=1 mm and y=-35 mm obtained with rhoCentralFoam (a), rhoPimpleFoam
(b), and experiments33 (symbols).
Table 2. Comparison of the frequencies (in Hz) predicted with OpenFOAM’s solvers and the code of Larcheveque et
al.2 The 1st mode in the experiments is at 1990 Hz.
12 of 21
tic engines and refrigerators are of practical interest because they have several advantages over their typical
counterparts, notably they are very simple and have at the most only one moving part.37 In spite of this, few
experiments and CFD simulations of validation quality (i.e., the test problem can be affordably reproduced
with CFD, there is enough information to reproduce it, and there is enough data for validation) are avail-
able for thermoacoustic engines,36, 38 a notable exception being Yu et. al.39 Therefore, the present section
only aims to compare qualitatively results from rhoPimpleFoam with the known physics of thermoacoustic
instabilities.40
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
Figure 12. Axisymmetric section of the thermo-acoustic engine considered in Sec. III.D. The heater is formed by 12
flat rings held at a constant temperature and located between x = 30 and x = 40. Dimensions are in mm.
The flow inside a thermoacoustic engine is modeled here using the Reynolds-Averaged Navier-Stokes
(RANS) conservation equations with the k − ǫ model with universal values for its constants and a turbulent
Prandtl number of 1. The computational domain considered is shown in Fig. 12 and corresponds to that
used by Zink et. al.36 It consists of a 2D axisymmetric section of a tube having a heater inside of it. This
heater is formed by 12 concentric flat rings 0.5 mm thick and 0.5 mm apart. The temperature of the rings is
held constant and varies linearly from 700 K at x=30 mm to 300 K at x=40 mm. This temperature gradient
is critical, as clarified later. All other boundaries are adiabatic. Walls are no-slip and are modeled with
low Reynolds number wall functions. The gas is air and modeled with a constant specific heat capacity of
1100 J/kg-K and temperature-dependent viscosities and conductivities according to Sutherland’s formula,
cf. Eq. 5. The mesh used for the geometry in Fig. 12 consists of 3318 predominantly hexahedral cells refined
near the heater plates. The minimum grid spacing is about 0.1 mm and the largest 1 mm. A 2nd order
time-advancement with a time step no smaller than 5e-9 s are used to run the simulations. Divergence terms
are discretized with a linear upwind scheme, gradients with a cellMDLimited scheme, and Laplacians with
linear limited schemes.
To illustrate how a steady flow inside the tube would look like, and also to follow the approach taken
by Zink et. al,36 the inlet boundary conditions is varied to produce a steady or unsteady operation. For a
steady operation the pressure is set to 101335 Pa at the inlet and 101325 at the outlet (using OpenFOAM’s
pressureInletOutletVelocity boundary condition). With these settings, rhoPimpleFoam produces a steady
rightward moving flow with a bulk velocity of 0.26 m/s and a temperature gradient near the heater as
depicted in Fig. 13. To produce the unsteady flow of main interest here, the inlet boundary condition is
abruptly changed from the one above to a wall, a change that can be interpreted as closing an inlet valve.
This change promotes the 1/4-wave-mode instability and concomitant oscillatory flow discussed next. (Such
a oscillatory flow was also exactly reproduced by using as an initial condition a quiescent flow with a uniform
temperature of 300 K.)
Figure 14 shows a pressure timetrace taken near the heater. Notice that, early on, pressure fluctuations
grow, and later saturate reaching a so-called limit cycle. The spectra corresponding to this pressure timetrace
is shown in Fig. 15 and exhibits a dominant mode at 635 Hz and two weaker ones at 1267 and 1899 Hz.
The dominant mode is close to that of 614 Hz corresponding to the 1/4-wave mode of the tube, which is
highlighted with a vertical line in Fig. 15. This frequency of 614 Hz is calculated with f = c0 /(4l), where
c0 is the sound speed evaluated at the mean temperature of the tube (368 m/s) and l is the tube length
(0.15 m). A dominant 1/4-wave mode is further supported by Fig. 16, which shows that the rms pressure
fluctuations are higher near the leftmost end of the tube, and decay in the axial direction reaching minimum
values at the rightmost.
When conducting CFD simulations, pressure timetraces such as that in Fig. 14 can be produced by
13 of 21
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
Figure 13. Temperature field (in K) and with superimposed unity velocity vectors at steady conditions.
5
x 10
1.25
1.2
1.15
1.1
1.05
p (Pa)
0.95
0.9
0.85
0.8
0 0.02 0.04 0.06 0.08 0.1 0.12
t (s)
Figure 14. Pressure timetrace taken near the heater showing the initial growth and later saturation of pressure
fluctuations.
14 of 21
12
11
10
9
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
PSD (au)
4
0 500 1000 1500 2000
f (Hz)
Figure 15. Power spectral density (PSD) of pressure fluctuations taken near the heater in logarithmic arbitrary units.
The vertical line denotes the frequency calculated for the 1/4-mode of the tube.
Figure 16. Field of the rms pressure fluctuations (in Pa) denoting a 1/4-wave mode.
15 of 21
numerical instabilities. Hence, care must be taken in distinguishing physical phenomena from numerical
artifacts. The most important piece of evidence supporting the fact that the behavior in Fig. 14 is not an
artifact are the in-phase pressure and heat release time traces shown in Fig. 17. Such an in-phase relationship
is a necessary condition for a thermoacoustic instability, as stated by Rayleigh’s criterion.40 In Fig. 17, the
time has been normalized with the acoustic period T of the dominant mode (0.00157 s), p∗ = (p − p0 )/prms ,
q∗ = (q − q0 )/qrms , p is pressure, q is the heat transfer from the heater to the fluid, the subindex 0 denotes
mean values, and the subindex rms denotes rms values.
1.5
1
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
0.5
p* or q*
−0.5
−1
−1.5
−2
64 65 66 67 68 69 70
t/T0
Figure 17. Temporal evolution of normalized pressure (thin lines), p∗ , and heat release (thick lines), q ∗ , taken near the
heater. Notice p∗ and q ∗ fluctuations are in phase with each other.
Further insight into the flow behavior during the limit cycle is given by Fig. 18. (In the following it is
useful to recall that the temperature in the heater varies from 700 K at x=30 mm to 300 K at x=40 mm.)
Near a pressure maximum, Fig. 18a, the flow is moving leftward and the high-temperature fluid is located
leftward of the heater. Since the plates of the heater are at a constant (time-independent) temperature, it
is at this point where the temperature gradient normal to the heater surface and, consequently, the heat
flux into the fluid, are maximum. As a result, pressure and heat-flux maximums coincide. Later on a flow
reversal occurs (at the left of the heater), Fig. 18b, and then the flow moves rightward. When a pressure
minimum is reached, Fig. 18c, the high-temperature fluid is now located near the heater plates, reducing
the heat flux according to the argument discussed above. Thus, pressure and heat flux minimums coincide.
Finally, after these minimums, the flow (at the right of the heater) reverses direction again, Fig. 18d, then the
high-temperature fluid is convected out of the heater region, and the cycle repeats. This sequence shows how
pressure and heat-fluxes are in phase from the movement of high-temperature fluid near the heater, due to
flow reversals that occur twice during an instability cycle. Moreover, this sequence shows that the “gas moves
hotward while the pressure is rising and coolward while the pressure is falling”.37 All these observations
agree qualitatively with the well-known operation of the Sondhauss tube as described by Rayleigh.37, 40 The
Sondhauss tube is somewhat similar to the thermoacoustic engine considered here but does not have a heater
inside the tube and maintains the temperature gradient at the tube (side) walls.
As seen in Fig. 14, there are very large pressure fluctuations, with an rms value of 11968 Pa, which is
orders of magnitude higher than those seen in practice. These high amplitudes were also observed in the
simulations of Zink et al.,36 and are consistent with the boundary conditions used in the simulations, which
do not allow for the acoustic losses through the boundaries happening in actual devices. One empirical way
to account for these losses is to use OpenFOAM’s nonreflective boundary condition, cf. Eq. 8, and choose a
value of l∞ that matches some experimental data. For the present problem, l∞ = 0.001 produces the same
high pressure fluctuations of Fig. 14, l∞ = 0.02 reduces them to a rms value of 6888 Pa, and l∞ = 0.1
kills the oscillations. Future work should develop models for acoustic losses, possibly through the use of
empirically-determined impedances, or consider a computational domain compose of the thermoacoustic
engine, as done above, plus an expansion at the rightmost end.
16 of 21
(a)
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
(b)
(c)
(d)
Figure 18. Temporal evolution of the instantaneous temperature field (in K) and unity velocity vectors during a
thermoacoustic cycle: (a) Near a pressure maximum. (b) Flow reversal between pressure extrema. (c) Near a pressure
minimum. (d) Flow reversal between pressure extrema.
17 of 21
IV. Summary, conclusions, and suggestions for future work
This paper evaluates the capability of the OpenFOAM solvers rhoCentralFoam and rhoPimpleFoam to
simulate acoustic and thermoacoustic phenomena in internal subsonic flows. To make this evaluation com-
prehensive, four different test cases, each of which highlights a particular phenomenon, have been considered.
In the problem of acoustic waves propagating with no mean flow inside a muffler, the above solvers predict
transmission losses profiles in good agreement with experiments; in particular, a high frequency peak of
practical interest is well predicted. For the case of an impulsively heated wall producing acoustic waves, the
OpenFOAM solvers tested in this work capture well the propagation of these waves inside a square cavity
when compared with a highly-accurate CFD code. Moving to a problem of hydrodynamic/acoustic interac-
tions, rhoCentralFoam and rhoPimpleFoam reproduce the Rossiter mechanism in cavity flows. According to
this mechanism, vortices impinge in the back side of the cavity producing acoustic waves, which in turn affect
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
the shear layers producing these vortices, closing a feedback loop. Moreover, these solvers predict both fre-
quencies and amplitudes of Rossiter modes agreeing well with data from experiments and another CFD code.
In the simulation of a thermoacoustic engine, rhoCentralFoam and rhoPimpleFoam qualitatively reproduce
the initial growth and later saturation of thermoacoustic instabilities, as well as their driving mechanism.
Future work should consider a quantitative assessment of these solvers to capture thermoacoustic instabilities
using, for example, the CFD and experimental data in Yu et. al.39
There are, nonetheless, some shortcomings with current OpenFOAM’s solvers (version 2.1.1) to sim-
ulate acoustic and thermoacoustic phenomena. First, the highly specialized boundary conditions needed
to simulate these phenomena, such as experimentally-determined complex impedances or highly-accurate
(nonperfectly) nonreflecting boundaries, are not currently available in the official version of OpenFOAM.
(In the muffler test problem, OpenFOAM’s nonreflecting boundary condition is found to work better when
combined with sponge zones.) Specialized boundary conditions are needed not only for the internal flows
considered here, but are critical to simulate external flows. Another shortcoming is the lack of an explicit
CFD solver, a type of solver that is better suited for the accurate computation of hyperbolic problems such
as acoustics. These shortcomings, nonetheless, can be surpassed since OpenFOAM provides all the elements
to develop both these specialized boundary conditions and new explicit CFD solvers.
Acknowledgements
The author would like to thank Dr. Michael Klassen for proofreading this paper and the financial
support from the Augmentor Group of the Arnold Engineering Development Complex of the U. S. Air
Force. This support comes from the SBIR project entitled ‘Computational Modeling of Coupled Acoustic
and Combustion Phenomena Inherent to Gas Turbine Engines’, contract number FA9101-13-M-0020.
18 of 21
Figure 19. Schematic of the Volvo test problem.41
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
tum, and total enthalpy. Turbulence closure is achieved with the one-equation LES model provided in
OpenFOAM with turbulent Prandtl and Schmidt numbers of 1. The mesh used for the present simulations
has a total of 345600 hexahedral cells, a minimum grid spacing of approximately 1.5 mm, and a maximum
of 4 mm. The simulations are run while keeping a maximum convective Courant number of 0.2, which is
defined as Co = ∆tuc /∆, with ∆t being the time step, ∆ a characteristic grid spacing, and uc a characteristic
convective speed. Hence, the resulting time steps might be to large to capture acoustic waves, which is fine
for what is intended here. Gradients and Laplacians are discretized with central schemes, while convective
fluxes are discretized with limited linear schemes, except the momentum fluxes which are discretized with a
so-called filtered linear scheme exclusively designed for LES.
Figures 20 and 21 show that simulations capture well the spatial variation of the mean horizontal velocity.
In particular, notice in Fig. 20 that the simulations predict well the recirculation length, i.e. the axial distance
between x=0 and the x location where the velocity changes sign. Moreover, the mean velocity profiles in
Fig. 21 are well captured. Note that the asymmetry of the experimental mean profile at x/h = 3.75 is an
artifact of the experiment,41 not a problem with the simulation results. Rms profiles are not captured as well
as mean profiles. Notice in Fig. 21 that the simulations show spikes in the rms profiles near the walls not
seen in the experiments. This should not be seen as a problem with the simulations because these spikes are
physical and known to exist in turbulent wall-bounded flows. Overall, the agreement between simulations
and experiments for the nonreacting case is satisfactory.
40
30
20
u (m/s)
10
−10
−20
0 0.1 0.2 0.3 0.4 0.5
x/h
Figure 20. Axial variation of the horizontal velocity at the center-plane (y=0) obtained from nonreacting simulations
(line) and experiments (symbols).41
References
1 Middelberg, J. M., Barber, T. J., Leong, S. S., Byrne, K. P., and Leonardi, E., “Computational fluid dynamics analysis
of the acoustic performance of various simple expansion chamber mufflers,” Proceedings of Acoustics, 2004, pp. 123–127.
2 Larchevêque, L., Sagaut, P., Mary, I., Labbé, O., and Comte, P., “Large-eddy simulation of a compressible flow past a
19 of 21
x/h=0.95 x/h=3.75
2 1.5
1 1
u/uinlet
0 0.5
−1 0
−1 0 1 −1 0 1
y/h
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
0.8 0.5
0.6 0.4
urms/uinlet
0.4 0.3
0.2 0.2
0 0.1
−1 0 1 −1 0 1
Figure 21. Vertical variation of the mean and rms horizontal velocity normalized by the inlet bulk velocity at different
x planes obtained from nonreacting simulations (line) and experiments (symbols).41
of buffeting noise,” Acta Acustica united with Acustica, Vol. 98, No. 4, 2012, pp. 600–610.
4 Weller, H. G., Tabor, G., Jasak, H., and Fureby, C., “A tensorial approach to computational continuum mechanics using
object-oriented techniques,” Computers in physics, Vol. 12, No. 6, 1998, pp. 620–631.
5 Lysenko, D. A., Ertesvåg, I. S., and Rian, K. E., “Modeling of turbulent separated flows using OpenFOAM,” Computers
OpenFOAM and FeatFlow,” International Journal of Computational Science and Engineering, Vol. 7, No. 3, 2012, pp. 253–266.
7 Wojciak, J., Schnepf, B., Indinger, T., and Adams, N., “Study on the Capability of an Open Source CFD Software for
Unsteady Vehicle Aerodynamics,” SAE International Journal of Commercial Vehicles, Vol. 5, No. 1, 2012, pp. 196–207.
8 Greenshields, C., Weller, H., Gasparini, L., and Reese, J., “Implementation of semi-discrete, non-staggered central schemes
in a collocated, polyhedral, finite volume framework, for high-speed viscous flows,” International Journal for Numerical Methods
in Fluids, Vol. 63, No. 1, 2010, pp. 1–21.
9 Deshpande, S. S., Anumolu, L., and Trujillo, M. F., “Evaluating the performance of the two-phase flow solver interFoam,”
OpenFOAM with LES,” International Journal of Computational Methods, Vol. 9, No. 03, 2012.
11 Mack, A. and Spruijt, M., “Validation of OpenFoam for heavy gas dispersion applications,” Journal of hazardous mate-
OpenFOAM,” V European conference on computational dynamics. ECCOMAS CFD, 2010, pp. 1–20.
14 Marzouk, O. A. and Huckaby, E. D., “A comparative study of eight finite-rate chemistry kinetics for CO/H2 combustion,”
Engineering applications of computational fluid mechanics, Vol. 4, No. 3, 2010, pp. 331–356.
15 Kassem, H. I., Saqr, K. M., Aly, H. S., Sies, M. M., and Wahid, M. A., “Implementation of the eddy dissipation model
of turbulent non-premixed combustion in OpenFOAM,” International Communications in Heat and Mass Transfer , Vol. 38,
No. 3, 2011, pp. 363–367.
16 Martin, S., de Ruijter, B., and Jemcov, A., “Modeling an enclosed, turbulent reacting methane jet with the premixed
simulation of Internal Combustion Engines,” International Multidimensional Engine Modeling Users Group Meeting, 2012.
20 of 21
18 Doolan, C. J., Coombs, J. L., Moreau, D. J., Zander, A. C., and Brooks, L. A., “Prediction of noise from a wing-in-junction
flow using computational fluid dynamics,” Proceedings of Acoustics 2012, Fremantle, Australia, 2012.
19 Ferziger, J. H. and Perić, M., Computational methods for fluid dynamics, Vol. 3, Springer Berlin, 1996.
20 Jasak, H., Error analysis and estimation for the finite volume method with applications to fluid flows, Ph.D. thesis,
diffusion equations,” Journal of Computational Physics, Vol. 160, No. 1, 2000, pp. 241–282.
24 Kurganov, A., Noelle, S., and Petrova, G., “Semidiscrete central-upwind schemes for hyperbolic conservation laws and
Hamilton–Jacobi equations,” SIAM Journal on Scientific Computing, Vol. 23, No. 3, 2001, pp. 707–740.
25 Selle, L., Nicoud, F., and Poinsot, T., “Actual impedance of nonreflecting boundary conditions: Implications for compu-
tation of resonators,” AIAA journal, Vol. 42, No. 5, 2004, pp. 958–964.
Downloaded by UNIVERSITY OF CAMBRIDGE on June 19, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-3960
26 Seybert, A. F. and Ross, D. F., “Experimental determination of acoustic properties using a two-microphone random-
excitation technique,” Journal of the Acoustical Society of America, Vol. 61, 1977, pp. 1362.
27 Poinsot, T. and Lele, S. K., “Boundary conditions for direct simulations of compressible viscous flows,” Journal of
of turbulent flows,” Physical Society of Japan, Journal, Vol. 54, 1985, pp. 2834–2839.
35 de Villiers, E., The potential of Large Eddy Simulations for the modeling of wall bounded flows, Ph.D. thesis, Imperial
College, 2006.
36 Zink, F., Vipperman, J., and Schaefer, L., “CFD simulation of a thermoacoustic engine with coiled resonator,” Interna-
tional Communications in Heat and Mass Transfer , Vol. 37, No. 3, 2010, pp. 226–229.
37 Backhaus, S. and Swift, G. W., “New varieties of thermoacoustic engines,” Proceedings of the Ninth International
Journal of Sound and Vibration, Vol. 227, No. 3, 1999, pp. 511–522.
39 Yu, G., Dai, W., and Luo, E., “CFD simulation of a 300Hz thermoacoustic standing wave engine,” Cryogenics, Vol. 50,
21 of 21