Вы находитесь на странице: 1из 13

Acoustic optimization for centrifugal fans

Esra Sorgvena) and Ylmaz Doanb)


(Received: 16 January 2012; Revised: 10 May 2012; Accepted: 10 May 2012)

The aim of this study is to perform a multi-objective optimization in order to reduce the noise level of a centrifugal fan in early design stages. The main goal is to
minimize the ow-induced noise while providing the required pressure rise at the
operating ow rate. The design procedure begins with a baseline fan. An optimization surface around the baseline design is formed by using the design of experiments method. Accordingly, geometric parameter sets and the corresponding
CAD models are generated. Flow through these fans is simulated via Reynolds
averaged Navier-Stokes equations. Flow data are evaluated to predict the noise
level and the fan performance. Neural network method is employed to determine
the family of optimum points. Two of the Pareto designs are manufactured with
rapid prototyping in order to validate the numerical optimization procedure
via aerodynamic and aeroacoustic experiments. Experiments show that both
designs satisfy optimization goals. 2012 Institute of Noise Control Engineering.
Primary subject classication: 11.4.2; Secondary subject classication: 75.3

INTRODUCTION

Fan noise is one of the major concerns in a large variety of industrial applications. It is essential to integrate
noise emission as a design criterion in the early design
stage in addition to aerodynamic performance and production cost in order to decrease the time and money
spent to design and produce fans.
In fan systems noise is generated aerodynamically,
mechanically and electromagnetically. This paper focuses only on the aerodynamically generated noise in
centrifugal fans, which can be classied as follows1:
1. Interaction between the rotor blades and distortions in the inow or stationary obstacles (like
diffuser blades, tongue etc.) (Fig. 1(a))
2. Relative circulation in blade passages (Fig. 1(b))
3. Boundary layer separation due to increased incidence at high ow rates or due to decreased incidence at low ow rates (Figs. 1(c) and (d))
4. Recirculation at the intake opening and in the
blade row (Fig. 1(e)).
The rst noise source arises periodically at the blade
passing frequency (BPF), which is called a tone. The
a)

b)

Yeditepe University Mechanical Engineering Department,


34755 Istanbul TURKEY; email: sorguven@yeditepe.
edu.tr.
Arcelik A.S., R&D Department, Vibration & Acoustic
Technologies, 34950 Istanbul TURKEY; email: yilmaz.
dogan@arcelik.com.

Noise Control Engr. J. 60 (4), July-August 2012

rest of the sources are non-periodic and provoked by


random force excitations. These contribute to the
broadband noise. In most industrial applications with
small centrifugal fans, both tonal and broadband components are perceived as annoying.
All components of the aerodynamically generated
noise can be predicted via computational aeroacoustics
(CAA). Modern CAA methods are based on the Lighthills analogy2,3, which is enhanced among others by
Curle4, Ffowcs Williams and Hawkings5, Farassat and
Myers6 and Brentner and Farassat7. An important problem arises in the aeroacoustic computation due to the
large difference between the aerodynamic and acoustic
scales. Acoustic waves inherit lesser amplitudes and
higher frequencies than the aerodynamic pressure
waves, which can be resolved with traditional computational uid dynamics (CFD) methods. The usual discretisation methods of CFD are too dissipative for the
computation of acoustic waves. A solution for this
problem is the hybrid CAA methods, which predict
ow and acoustic elds separately (Fig. 2). In the rst
step the unsteady turbulent ow eld is simulated to determine the acoustic sources. In the second step the
propagation of the acoustic sources in near and far
acoustic elds is computed. In the literature, a great
number of models exist to solve the wave propagation.
Most of these models are based on linearizing the
governing uid dynamics equations. Some of these
methods are uid-acoustic decomposition811, acoustic
perturbation equations (APE)12,13, acoustic-viscous
splitting14,15 and expansion about incompressible
379

Fig. 1Noise generation mechanisms in a centrifugal fan.


formulation (EIF)16,17. The accuracy of noise prediction
depends basically on the modeling errors in ow
simulation, where the acoustic sources are computed.
Sophisticated ow simulations, which involve less
modeling errors, result in accurate acoustic source
terms; however, they increase the computational cost
drastically. For example a large eddy simulation
(LES) based noise source prediction for a centrifugal
fan may take months of pure computation time with
state-of-art parallel computers18. Such simulations can
only be performed in order to gain physical insight
and understand general noise generation mechanisms.
In order to implement noise prediction into early design stage, one needs to shorten the computational
requirements.
During the design stage usually several geometries
are proposed and these are compared to each other to
determine a geometry which satises the design criteria. Hence, during the early design stage it is enough
to determine the design with the lowest noise emission
rather than to determine the exact noise level of each
proposed design. Although accurate noise prediction

is a time-consuming process, which requires accurate


ow simulation and acoustic computation, comparison
of noise levels of different geometries can be performed
within reasonable time requirements with less sophisticated computational methods.
This paper proposes a multi-objective aeroacoustic
optimization method for fans, which can be easily performed in early design stages. Multi-objective optimization problems have been studied over three decades
using diverse techniques and providing successful solutions to a variety of problems1921. However, it has not
been applied to noise reduction yet.
The design procedure begins with an initial baseline
design. The baseline can be a new fan designed according to the traditional turbomachinery design correlations, or an existing fan, which is to be improved.
With the design of experiments (DoE) method, geometric parameters of the baseline are varied systematically.
Reynolds averaged Navier-Stokes (RANS) simulations
are performed for each of the proposed fan geometries.
Flow data, which provide an indication on aerodynamic
and acoustic performance, are extracted from the

Fig. 2An overview of modern CAA methods ( represents any ow variable and rac is the
acoustic pressure).
380

Noise Control Engr. J. 60 (4), July-August 2012

simulations. The extracted ow data are modeled as


functions of the geometric parameters. These functions
are minimized or maximized to achieve the optimization goals. Finally, a family of fan geometries, which
provide the best aerodynamic and aeroacoustic properties is determined. The proposed method is demonstrated for a backward curved centrifugal fan. With
this strategy, sound power level is reduced by 6 dB on
the average over the whole operation range.

ACOUSTIC PREDICTION BASED ON


FLOW DATA

The direct numerical simulation (DNS) is the most


accurate ow prediction method, since it solves the
Navier-Stokes equations without any modeling. The
whole range of time and length scales (from integral
scales to Kolmogorov scales) is resolved. Thus, the
acoustic scales are directly resolved, i.e. an additional
computation of acoustic sources and wave propagation
is redundant (Fig. 2). Unfortunately, it is only applied
in low Reynolds number ows because of the practically
unaffordable discretisation requirements (numerical

effort of DNS-Re4). Large eddy simulation (LES) is


the second most accurate method, which resolves the
large and energetic scales of turbulence and models
the small and dissipative scales. Although computational requirement of LES is considerably lower than
DNS, it is still high for industrial applications. Reynolds
averaged Navier Stokes (RANS) method is based on averaging the ow variables over a large time range.
RANS is the fastest ow simulation method; however,
it fails in accuracy mainly due to the deterministic approach of its formulation. The decomposition of the turbulent energy spectrum with different turbulence
models is symbolically represented in Fig. 3. RANS
results do not contain any information on acoustic
waves because of the inherent modeling errors. Still,
any other method cannot be chosen in the design stage
because of the time limitations. One approach to overcome this deciency is to employ empirical methods,
which estimate turbulent uctuations based on the averaged ow eld2224. These stochastic methods can be
combined with a RANS solution in order to estimate unsteady turbulent uctuations, which can then be used to
compute acoustic sources generated in the ow eld.

Fig. 3Kinetic energy decomposition in the presence of a predominant frequency, top: with DNS,
middle: with LES, bottom: with unsteady RANS formalismus.
Noise Control Engr. J. 60 (4), July-August 2012

381

Fig. 4Instantaneous vorticity distributions a) Fan I, b) Fan II.


Although there are well evolved models for stationary
blunt bodies such as airfoils25, a stochastic model for rotating blades does not exist.
The approach followed in this study is to employ
RANS simulations to compare the overall sound power
level (OSPL) of different fan geometries, rather than to
predict the exact noise level of each fan. For this purpose, a preliminary investigation is performed in order
to determine the ow variables, which can be computed
with RANS simulations and also provide a good indication for the OSPL of a fan. In a preliminary study18,
two fans are investigated via aerodynamic and acoustic
experiments, as well as via LES and RANS. Both fans
are centrifugal fans with similar dimensions, performance curves and operating points. Nevertheless, one
of the fans has much lower OSPL than the other. The
aerodynamic and acoustic elds are compared to determine a correlation between the powerful acoustic
sources and ow variables. Two ow variables, vorticity and pressure distribution are found to be strongly
correlated with the acoustic properties.
Results of the preliminary study are summarized in
Figs. 4 and 5 to demonstrate the effect of vorticity and

pressure distribution on the OSPL. Figure 4 shows the


instantaneous distribution of the vorticity magnitude
extracted from LES on a cross-section across the impeller for both fans. In both fans vorticity is produced
mainly around the blades, especially on the trailing
edge. However, in the louder fan vorticity is not dissipated immediately, but transported further with the ow
through the volute. Furthermore, large vorticity magnitudes indicate a strong tongueblade interaction and
poor ow conditions through the volute. Hot spots upstream the blades indicate that inow is not axisymmetric, which causes an unbalanced loading on the blades.
Comparison of LES results shows that large vorticity
magnitudes indicate large OSPL values. Therefore the
average vorticity magnitude on a cross-section across
the impeller is chosen as an indicator for OSPL.
Figure 5 shows the instantaneous pressure distribution. Pressure contours exhibit a high spatial gradient in
the neighborhood of the tongue. In both fans, as the radius of the volute increases, the gap between the blades
and the volute increases and the effect of the pressure
waves due to blade motion on the volute surface
decreases. Spatial pressure gradients on the volute

Fig. 5Instantaneous pressure distributions a) Fan I, b) Fan II.


382

Noise Control Engr. J. 60 (4), July-August 2012

Fig. 6Schematic view of the baseline fan, left: side view, right: cross-sectional view.
decrease in rotational direction. In the louder fan, magnitude of the pressure gradients on the volute is higher. To
quantify this phenomenon, the standard deviation of the
pressure on the casing, pstd, is computed.
!
N
1 X
2
pstd
1
pi  p
N  1 i1
where N is the total number of nodes, pi is the pressure at
the node i and 
p is the surface-averaged pressure value.
Large values of pstd indicate powerful sound sources
and a high OSPL.
In summary, two variables are chosen as indicators of
the OSPL: the average vorticity on a cross-section
across the impeller and the standard deviation of pressure on the volute. Since RANS provides a fair prediction for average ow variables, both of these ow
variables can be predicted accurately with RANS
simulations.

DESIGN OF EXPERIMENTS

A centrifugal fan, which is shown in Fig. 6, is


selected as the baseline. Reynolds number based on
the blade tip diameter is Retip = 1.7106 and Mach
number at the tip is Matip = 0.06.
Aerodynamics and aeroacoustics of a fan is affected
by a number of geometrical parameters, such as the
shape of the volute, inlet and exit areas, number of
blades, blade thickness, length, height and angles. In
order to keep the calculations simple and focus on the
optimization strategy, the effect of only four geometric
variables on the OSPL is investigated. These are the
blade inlet and outlet angles (b1 and b2) and the blade
inlet and outlet radii (r1 and r2).
With the help of the DoE method the number of
combinations to be tested is reduced. The DoE
method provides a well-distributed set of parameters,
Noise Control Engr. J. 60 (4), July-August 2012

which can be used for surface tting. As a result,


24 combinations of the investigated parameters are
determined (Table 1) and corresponding impeller geometries are generated. The new impellers have all
other geometric variables (such as volute shape, blade
height, thickness etc.) the same as the baseline
design.

Table 1Geometric variables of the simulated


fans.
Number of
geometries

r1 (mm)

r2 (mm)

b 1 ( )

b 2 ( )

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24

32
32
42
42
37
37
37
37
32
32
42
42
37
37
37
37
32
32
42
42
37
37
37
37

63
67
63
67
65
65
65
65
65
65
65
65
63
63
67
67
65
65
65
65
63
63
67
67

85
85
85
85
80
80
90
90
85
85
85
85
80
90
80
90
80
90
80
90
85
85
85
85

38
38
38
38
33
43
33
43
33
43
33
43
38
38
38
38
38
38
38
38
33
43
33
43

383

averaged values are denoted with an overbar. The closure problem, which arises due to the emergence of
the Reynolds stress tensor, is solved with the Boussinesq approximation by dening the eddy viscosity mt.


@ui @uj
2
0
0
rui uj mt
dij rk

3
@xj @xi
3
The k-e turbulence model calculates the eddy viscosity as a function of the turbulent kinetic energy k and the
turbulent dissipation e.
mt rCm

4.1

@ rkui
@

@xi
@xj

COMPUTATIONAL METHODS

Steady RANS simulations are performed for all


designs with the commercial software Fluent26. The
continuity and the RANS equations are given below.





uj
@ r
ui 
@p @tij @ rui 0 uj 0



rgj
@xj @xi
@xi
@xi

2
3

Here, ui is the ow velocity in i direction, p is pressure, tij is the shear stress, g is the gravitational acceleration and rui 0 uj 0 is the Reynolds stress tensor. The time

Fig. 8Vorticity distributions.


384

Noise Control Engr. J. 60 (4), July-August 2012



 
mt @k
m
sk @xj

Gk Gb  re  Ym Sk

Computational Fluid Dynamics

@ui
0
@xi

Here r is density and Cm is a constant with the value


0.09. The transport of k and e is calculated with the
following partial differential equations.

Fig. 7Computational mesh.


4

k2
e

@ reui
@

@xi
@xj



 
mt @e
m
se @xj

e
e2
C1e Gk C3e Gb  C2e r Se
k
k

Where Gk is the generation of turbulence kinetic energy due to the mean velocity gradients (Gk = mt 2jSij Sijj,
where Sij is the strain tensor), Gb is the generation of turbulence kinetic energy due to buoyancy, Ym is the contribution of the uctuating dilatation in compressible
turbulence to the overall dissipation rate. C1e, C2e and
C3e are constants and sk and se are the turbulent Prandtl

Fig. 9Streamlines of relative velocities.


numbers for k and e, respectively. Sk and Se are source
terms. In this study, since the ow through the fan
is isothermal and incompressible, the terms Gb, Ym,
Sk and Se are zero. The values of the model constants
are: C1e = 1.44, C2e = 1.92, sk = 1.0 and se = 1.3.
In order to assure grid-independent results, meshing
is performed in two steps. First the coarsest mesh
resulting grid-independent ow data is determined by
repeating the ow simulation of the baseline design
with meshes of varying renement levels, and comparing their results. At the second step all CAD models are
meshed systematically with the predetermined meshing
parameters. Parameters that are kept constant for each
computational grid are the normal wall distance (y +),
growth rate, boundary layer thickness and mesh sizes
on the boundaries. The volume mesh around the impeller is then generated with hexahedral control volumes
(Fig. 7). The stationary part is meshed with unstructured tetrahedral cells. As a result of this procedure,
regions where large derivatives occur (e.g. boundary
layer) are represented with similar meshes. However,
some variations in the total number of control volumes
occur. The mean number of control volumes is 2106.
The computational domain is divided into two zones,
one surrounding the rotating impeller and the other surrounding the stationary volute. Zones are coupled via a
sliding interface and mass balance is forced across the
sliding interface. In order to minimize interpolation
errors, the ratio of the control volumes across the sliding interface is kept below 4:1. The employed boundary
conditions are no-slip on the walls, constant mass ow
rate on the inlet and constant static pressure on the outlet. In all simulations the ow rate and the rotational
speed are kept constant at 65 l/s and 2900 rpm, respectively. Spatial discretisation is performed with 2nd order central differencing scheme.
Noise Control Engr. J. 60 (4), July-August 2012

4.2

Acoustic Optimization

Optimization is performed to enhance the aeroacoustic properties of the baseline fan, i.e. to decrease its
OSPL. For this purpose two optimization objectives
are formulated:

Minimize the average vorticity on the crosssection across the impeller.


Minimize the standard deviation of pressure on
the volute.

An important point is that the aerodynamic performance should not be affected adversely by lowering
the OSPL. A fan performance curve represents the
pressure rise developed versus ow rate at a constant
rotational speed. In order to predict the performance
curve numerically, several CFD runs at varying ow
rates have to be performed for each fan. However, the
number of simulations required for a successful optimization process is already so large at to forbid an extensive numerical study. Considering the required
computation time and the data processing effort, ow
is simulated only at the operating point for each fan. It
is assumed that the performance curve is characterized
by the pressure rise developed at that ow rate. In order
to guarantee that the required aerodynamic performance

Table 2Correlations between the ow and geometrical variables.


P
oave
pstd

r1

r2

b1

b2

0.252
 0.192
0.224

0.823
0.237
0.757

0.008
 0.106
 0.044

 0.260
 0.451
 0.312

385

Fig. 10Response surface for the pressure rise P.

Fig. 11Response surface for the average vorticity oave.


386

Noise Control Engr. J. 60 (4), July-August 2012

Fig. 12Response surface for the standard deviation of pressure pstd.

is achieved, the minimum pressure rise that the fan has


to provide at the operating point is set as an optimization
constraint:

At the operating point (i.e. 65 l/s at 2900 rpm)


the pressure rise should be at least 160 Pa.

Response surface methodology is used to explore the


relationships between the extracted ow variables and
the geometric parameters. A surface is interpolated to
represent each of the ow variables as a function of
b1, b2, r1 and r2. Since the relationship between the
explanatory and response variables are complicated,
simple techniques such as single value decomposition
or polynomial responses did not provide satisfactory
results. Neural network method is used to generate the
response surfaces. The present optimization requires
simultaneous minimization of two ow variables. In
contrast to single objective optimization problems, the
solution of this multi-objective optimization is not a
single point, but a family of points referred to in the literature as Pareto points or Pareto front27. An example
to explain the theory of Pareto front is the 1-variable
2-objectives problem used in Ref. 28. Objectives of this
problem are:
Noise Control Engr. J. 60 (4), July-August 2012

Minimize f11 = x2
Minimize f12 = (x  2)2

Any point between 0 and 2 can be a solution for this


problem. So, here the Pareto front is {xj0 x 2}.

NUMERICAL RESULTS

One of the optimization objectives is to minimize


vorticity. Simulation results show that the generation
and transport of vorticity vary greatly with the geometric parameters of the impeller. Figure 8 presents the
comparison of vorticity distributions in two of the investigated impeller geometries. In the impeller #22,
vorticity generation is limited in the near vicinity of
the blades. However, in the impeller #23, even at the inlet large vorticity magnitudes are visible. The blade geometry causes an unbalanced blade load, which triggers
reverse ows and vorticity generation.
Figure 9 shows the comparison of streamlines of relative velocity in two impeller geometries. Reverse ow
through the blades in the upwind of the tongue is not
uncommon in centrifugal fans. Here, such a reverse
ow region is visible in all of the impeller geometries.
However, the extent of the reverse ow varies. Since
387

Fig. 13Relative velocity vectors and vorticity distribution of the rst Pareto design.
the reverse ow region obstructs the outlet area, magnitude of the outlet velocity increases locally. This causes
an increase in the OSPL and also a decrease in the aerodynamic performance. In an optimized fan, reverse ow
region should be minimal.
Pressure rise, average vorticity and standard deviation of pressure are extracted from RANS simulations.
The correlations between the ow and the geometric
variables are summarized in Table 2. The blade outlet
radius r2 has a large effect on all ow variables. The
correlation between r2 and P is the largest, which is
a direct result of the Eulers turbomachinery equation:
w U2 c2  U1 c1

where w is the work done on the uid per unit mass, U


is the blade circumferential speed and c is the absolute
tangential ow velocity and 1 and 2 represent the blade
inlet and outlet, respectively. Work is pressure rise, p,
over density (w = p/r). Eulers turbomachinery equation assumes the following form, if it is applied to a
centrifugal fan with no whirl at the inlet (c1 = 0):
pideal rr
 2 c2

2
r U2  U2



Q
cot b2 ;
2pr2 b2

where pideal is the ideal pressure rise, is the angular


velocity ( = pn/30, where n is the rotational speed in

Fig. 14Relative velocity vectors and vorticity distribution of the second Pareto design.
388

Noise Control Engr. J. 60 (4), July-August 2012

Fig. 15The experimental setup.


min-1), Q is the volumetric ow rate and b2 is the rotor
blade outlet width. In the ideal case the correlation between the pressure rise and the outlet blade radius r2 is
1, since these are in a linear relationship. Here, the correlation decreases to 0.82 due to losses, reverse ows
and leakage.
Second important geometric variable is the outlet
blade angle b2. Negative correlations show that an
increase in b2 decreases the listed ow variables
(Eqn. (8)). The average vorticity, which is the main
indicator of OSPL is mostly affected by b2.
The inlet blade angle and radius have only a limited
inuence on the ow variables. Since there is no prewhirl (c1 = 0), the second term in Eqn. (7) disappears
and the dependence of pressure rise on inlet blade angle
and radius vanishes.
The data extracted from RANS are fed into the optimization procedure, where the neural network method
is employed to t a surface for each ow variable as a
function of the geometric variables. The response surfaces are created in a 5-dimensional space, where each
ow variable is represented as a function of r1, r2, b1
and b2. 3D and 2D projections of the tted response

surfaces are shown in Figs. 10 through 12. In the projections x- and y-axes are chosen as r2 and b2, since
these geometric parameters have the largest effect on
the investigated ow variables.
The objectives of the optimization procedure are to
minimize both pstd and oave. As the response surfaces
demonstrate, these objectives are competitive. The local
minimum and maximum of pstd and oave surfaces are at
different locations. Furthermore the optimization constraint (i.e. p should be at least 160 Pa) eliminates
a number of points in the optimization space, which
would lead to low pstd and oave values. With these goals
and the constraint, optimization procedure results in
several Pareto designs. Two of the Pareto designs,
which have the longest and the smallest blade chord
lengths, are chosen for further study. These designs
are investigated both numerically and experimentally.
Both Pareto designs are manufactured via rapid prototyping and tested to determine the OSPL values.
Average vorticity distribution and the relative velocity vectors of the Pareto designs are presented in
Figs. 13 and 14. In both Pareto designs no reverse ow
through the blades is visible and large vorticity magnitudes are limited in the near vicinity of the blades.

EXPERIMENTAL VALIDATION

Prototypes of both Pareto designs are manufactured


with rapid prototyping and OSPL values are measured
over the whole operating range. Acoustic measurements
are performed in a semi-anechoic room, which has a
background noise level of less than 20 dB. Dimensions
of the semi-anechoic room are 4  4  5 m, so that
the minimum measurable frequency is 125 Hz, which
is below the frequency range of interest in this study.
Experiments are performed according to the ANSI
standards29. In the experimental setup (Fig. 15) mylar,
an acoustically transparent material, is used to obtain a

Fig. 16OSPL of the baseline and Pareto designs with respect to pressure rise.
Noise Control Engr. J. 60 (4), July-August 2012

389

control volume for mass ow and a free eld boundary


for acoustic waves. The maximum limits of the experimental setup are 1 m3/s and 750 Pa for ow rate and
pressure rise, respectively. OSPL is measured with respect to pressure rise, which is chosen as the control
variable in the experiments. Pressure rise through the
fan is controlled by adjusting the air outow area. A
differential pressure sensor with  1% accuracy is used
to measure the pressure rise. Reproducibility and repeatability of the experiments are ensured before evaluating the experimental results. Figure 16 shows the
comparison of OSPL curves of the baseline and Pareto
designs. Both Pareto designs are acoustically better
than the baseline.

2.
3.
4.
5.

6.
7.

8.

CONCLUSION
9.

In the frame of this work, a multi-objective acoustic


optimization of a centrifugal fan is performed. The objective functions are identied by comparing detailed
ow data obtained via LES with acoustic measurements
and RANS simulations. With this strategy we found
that for small centrifugal fans pstd and oave should be
minimized to minimize the OSPL. These objective
functions should be valid for any fan having low pressure rise and low ow rate. This optimization strategy
can be applied to axial, centrifugal and mixed type fans.
But, if the operating conditions change dramatically, so
that the noise generation mechanisms change, then the
objective functions may not be valid any more. Nevertheless, the main strategy, i.e. comparing detailed ow
data with OSPL and identifying ow variables, which
inuence OSPL and trying to optimize these ow variables, can help to reduce ow-induced noise in a variety
of applications including but not limited to fans.
To evaluate the acoustic performance of the optimized
fans the OSPL of a set of commercially available fans
with similar pressure rise and ow rate is measured with
the same experimental setup and procedure. At the operating point the OSPL values for the commercial fans are
measured between 76 and 81 dB. At the operating point
OSPL of the optimized fans is 74.5 and 76.5 dB.
The sole input of the optimization procedure is the
ow data extracted from steady RANS simulations.
Since hardware and computation time requirements
for steady RANS simulations are low, optimization
can be performed with reasonable resources. This
makes the integration of acoustic optimization into the
early design stage possible.

10.
11.
12.
13.
14.
15.
16.

17.

18.
19.
20.
21.

22.

REFERENCES

23.

1.

M. Cudina and J. Prezelj, Noise generation by vacuum cleaner


suction units Part I. Noise generating mechanisms An overview, Applied Acoustics, 68, 491502, (2007).

24.

390

Noise Control Engr. J. 60 (4), July-August 2012

J.M. Lighthill, On sound generated aerodynamically. I. General


theory, Royal Society of London (A), 211, 564587, (1952).
J.M. Lighthill, On sound generated aerodynamically. II. Turbulence as a source of sound Royal Society of London (A),
222, 132, (1954).
N. Curle, The inuence of solid boundaries upon aerodynamic
sound, Royal Society of London (A), 231, 505514, (1955).
J.E. Ffowcs Williams and D.L. Hawkings, Sound generation
by turbulence and surfaces in arbitrary motion, Philosophical
Transactions of the Royal Society (A), 264, 1151, 321342,
(1969).
F. Farassat and M.K. Myers, Extension of Kirchhoffs formula
to radiation from moving surfaces, J. Sound and Vibr., 123(3),
451460, (1988).
K.S. Brentner and F. Farassat, An analytical comparison of the
acoustic analogy and Kirchhoff formulation for moving surfaces, American Helicopter Society 53rd Annual Forum,
13791386, (1997).
C. Bailly and D. Juve, Numerical solution of acoustic propagation problems using linearized Euler equations, AIAA Journal,
38(1), 2229, (2000).
C. Bogey, C. Bailly and D. Juve, Computation of ow noise
using source terms in linearized Euler equations, AIAA Journal,
40(2), 235243, (2002).
K. Wisvanathan and L.N. Sankar, Toward the direct calculation
of noise: Fluid/Acoustic coupled simulation, AIAA Journal, 33
(12), 22712279, (1995).
R.R. Mankbadi, R. Hixon, S.H. Shih and L.A. Povinelli, Use
of linearized Euler equations for supersonic jet noise prediction, AIAA Journal, 36(2), 140147, (1998).
R. Ewert, M. Meinke and W. Schrder, Computation of trailing edge noise via LES and acoustic perturbation equations,
AIAA Paper 2002-2467, (2002).
R. Ewert and W. Schrder, Acoustic perturbation equations
based on ow decomposition via source ltering, Journal of
Computational Physics, 188, 365398, (2003).
J.C. Hardin and D.S. Pope, An acoustic/viscous splitting technique for computational aeroacoustics, Theoretical and Computational Fluid Dynamics, 6, 323340, (1994).
J.A. Ekaterinaris, New formulation of Hardin Pope equations for aeroacoustics, AIAA Journal, 37(9), 10331039,
(1999).
S.A. Slimon, M.C. Soteriou and D.W. Davis, Computational
Aeroacoustics simulations using expansion about incompressible ow approach, AIAA Journal, 37(4), 409416,
(1999).
S.A. Slimon, M.C. Soteriou and D.W. Davis, Development of
computational aeroacoustics equations for subsonic ows using
a Mach number expansion approach, Journal of Computational Physics, 159, 377406, (2000).
E. Sorguven, Y. Dogan, F. Bayraktar and K.Y. Sanliturk, Noise
prediction via large eddy simulation: Application to radial
fans, Noise Control Engr. J., 57(3), 169178, (2009).
K. Deb, Multi-Objective Optimization using Evolutionary
Algorithms, John Wiley and Sons, New York, (2009).
C.A.C. Coella and G.B. Lamont, Applications of MultiObjective Evolutionary Algorithms, World Scientic Pub.,
London, (2004).
S. Derakhshan, B. Mohammadi and A. Nourbakhsh, The comparison of incomplete sensitivities and Genetic algorithms
applications in 3D radial turbomachinery blade optimization,
Computers and Fluids, 39(10), 20222029, (2010).
W. Bchara, C. Bailly and P. Lafon, Stochastic approach to
noise modeling for free turbulent ows, AIAA Journal, 32(3),
(1994).
M. Billson, L.E. Eriksson and L. Davidson, Jet noise prediction using stochastic turbulence modeling, 9th AIAA/CEAS
Aeroacoustic Conference, (2003).
J. Ostertag, G. Guidati, S. Guidati and S. Wagner, Validation
of an airframe noise prediction methodology based on BEM

and statistical turbulence modeling, 2nd Aeroacoustics Workshop, SWING, (2000).


25. R. Ewert, J.W. Delfs and M. Lummer, The simulation of
airframe noise applying Euler-perturbation and acoustic
analogy approaches, Int. J. of Aeroacoustics, 4, 6991,
(2005).
26. Fluent Version 6.2, Users Guide, USA, (2005).

Noise Control Engr. J. 60 (4), July-August 2012

27. P. Sen and J.B. Yang, Multiple Criteria Decision Support in


Engineering Design, Springer, London, (1998).
28. T.L. Vincent and G.J. Grantham, Optimality in Parametric
Systems, John Wiley and Sons, New York, (1981).
29. ANSI S12.11-2003 / Part 1, American National Standard,
Acoustics, Measurement of noise and vibration of small air
moving devices, Part 1: Airborne noise emission.

391

Вам также может понравиться