Você está na página 1de 99

AIR MOVEMENT AND CONTROL

ASSOCIATION INTERNATIONAL, INC.


The International Authority on Air System Components

FA N A C O U S T I C S
Noise Generation and Control Methods

Alain Guédel C E T I AT
Fan Acoustics
Noise Generation and Control Methods

Alain Guédel

© 2007 by Air Movement and Control Association International, Inc.

All rights reserved. Reproduction or translation of any part of this work beyond that permitted by Sections
107 and 108 of the United States Copyright Act without the permission of the copyright owner is unlawful.
Requests for permission or further information should be addressed to the Executive Director, Air
Movement and Control Association International, Inc. at 30 West University Drive, Arlington Heights, IL
60004-1893 U.S.A.
TABLE OF CONTENTS
Chapter 1 - Fan Noise

1.1 The different types of fans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1

1.2 General remarks on fan noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5

1.3 Parameters of influence, modeling, and control of fan noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .25

1.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .60

Chapter 2 - Fan Noise Installation Effects

2.1 Definitions and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .61

2.2 Effect of inflow conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .65

2.3 Acoustic loading effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72

2.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .78

Annex A - Fan Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .79

Annex B - Acoustics Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .82

Annex C - Forward Curved Centrifugal Fan Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .86

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .88
PREFACE
This book fills a much needed gap in the literature of ventilation engineering. There have been many
volumes published concerning fan types, similarity rules, ducting system calculations and design,
applicational problems, etc. At the same time, there has been a paucity of published information on the
acoustics and noise control associated with these fans and their systems. Many engineers are wary of
entering this “black art,” as it is perceived by the non-acoustician. To lead you through the veritable maze
of half-truths and empiricism which surrounds the subject, there is none better equipped than Alain Guédel.
I have known Alain for a number of years through association on ISO, CEN and industry committees, and
have valued his friendship and hard work. He had contributed greatly to the production of many documents
and has assisted our endeavors by carrying out technical experiments in support of our deliberations. His
knowledge of the science and the practice of fan acoustics is exceptional so that the reader may use this
text with confidence.

W.T.W. Cory
Convenor ISO TC117/WG2
1 Fan Noise

1.1 The Different Types of Fans

1.1.1 Definition of a fan

ISO Technical Committee 117 defines a fan as "a rotary bladed machine that receives mechanical
energy and utilizes it by means of one or more impellers fitted with blades to maintain a continuous
flow of air or other gas passing through it, and whose work per unit mass does not normally exceed
25,000 J/kg" (1996).

The machine is no longer considered a fan if the work per unit mass exceeds a value of 25,000
J/kg. It is then referred to as a turbo-compressor.

There are four main categories of fans, which are introduced in the following sections and
discussed at length in Section 1.3. Other types of fans do not appear in the description below as
they are rarely used in common practice (annular fans, for example).

1.1.2 Axial-flow fans

Flow through axial-flow fans is essentially parallel to the impeller axis (Figure 1). The geometry of
the blades and of the shroud depends on the work provided by the impeller. Axial-flow fans are
usually used to convey large volume flow against low or medium pressure. There are three types
of axial-flow fans: vaneaxial, tubeaxial, and propeller (ISO Technical Committee 117, 1996). These
fans will be described in Section 1.3.1.

Figure 1
Axial Flow Fan

1
Fan Noise

1.1.3 Centrifugal fans

Flow enters the impeller axially and leaves at its periphery in the radial direction (Figure 2). There
are three types of centrifugal fans, determined by their blade shape.

-backward curved blades


-radial blades
-forward curved blades (or squirrel cage centrifugal fans)

Due to the centrifugal effect, which supplements the aerodynamic work of the blades, centrifugal
fans usually convey moderate flow at high pressure, especially backward curved fans.

Figure 2
Centrifugal Fan

1.1.4 Mixed flow fans

In these fans, the flow enters axially and leaves the impeller with an angle of 30 to 80 degrees from
the axis (Figure 3). The fan pressure rise is due to the combined effects of blade work and
centrifugal effect.

Figure 3
Mixed Flow Fan

2
The Different Types of Fans

1.1.5 Cross-flow fans

The impeller of the cross-flow fan is very similar to that of a forward curved centrifugal fan, but the
passage of the flow through the impeller is very different as the flow is essentially 2D and remains
in a section normal to the axis. The fan operation depends on a vortex that is located across the
blades near the tongue (Figure 4). The position of the vortex varies with the fan operating point.

Cross-flow fans are used in systems in which substantial flow at low pressure are required. The
flow volume is proportional to the impeller width.

Figure 4
Cross-Flow Fan
Tongue

1.1.6 Fan performance

The fan performance curve is a plot of fan pressure rise against volume flow. Figure 5 compares
the performance curves of several types of axial and centrifugal fans (Daly, 1992). Each of these
fans has a constant air power of 10 kW at its peak efficiency point (see Annex A). This figure
exhibits the large differences between the curves, and it also shows differences in impeller
diameter and rotational speed, according to the fan.

The choice of a fan for a given application depends on parameters such as operating point,
available space, electric motor, cost, sound level, mechanical constraint, etc.

It is seldom easy to replace a fan of one type with a fan of another type. Conversely, fans of the
same type may have very different aerodynamic and/or acoustic performance.

3
Fan Noise

Pa

5000

A
4500

B
4000 ∆

C
Fan Total Pressure, Pt (Pa)

3500

D
3000

2500 E

2000 F

Ø
1500 ∆
G
Ø

1000 ∆
Ø

Ø

500
Ø
Ø ∆

5 10 15 20 25 30 35
Airflow, Q1 (m3/s)
Each fan, operating at top speed and best efficiency point Δ is chosen for an output
Q х P1 = 10 kW
Peak input power is taken at Ø
Drawings are to a uniform scale of 1:120

A: Backward-curved centrifugal fan - 42 rev/s E: Forward-curved centrifugal fan - 18 rev/s


B: Backward-curved centrifugal fan - 36 rev/s F: Axial-flow fan of hub ratio 0.35 - 24 rev/s
C: Axial-flow fan of hub ratio 0.5 - 48 rev/s G: Axial-flow fan of hub ratio 0.25 - 12 rev/s
D: Forward-curved centrifugal fan - 18 rev/s

Figure 5 - Comparative Performances of Fans of Different Types

4
General Remarks on Fan Noise

1.2 General Remarks on Fan Noise

1.2.1 Origins of fan noise

1.2.1.1 The three noise sources

Fan noise results from aerodynamic, electromagnetic and mechanical noise mechanisms.

Aerodynamic noise, due to the interaction of the flow with the rotating and fixed components of
the fan, is the source that is dealt with throughout this book as it is more prominent than the two
other sources for most fans. This document describes the main origins of the aerodynamic noise
in the next section, and in more detail according to each fan type, in Section 1.3.

Electromagnetic noise refers to the noise of the electric motor, which drives the fan. The motor
is most often an induction squirrel-cage motor of single or three-phase type according to the fan
power. Motor noise emanates from the bearings, the noise of the cooling fan, and the periodic
forces due to the magnetic field between the rotor and the stator (Cory 1992, 305-328; Guedel and
Mairet 1990). These fluctuating forces induce vibrations in the stator and the motor casing; the
latter radiating airborne noise to the outside. These vibrations may also generate structural noise
when they are transmitted to the fan casing through the motor supports.

The frequency spectrum of electromagnetic noise consists of peaks at multiples of the shaft
rotation frequency, N, especially at frequency R × N, where R is the number of rotor slots, and at
harmonics of twice line frequency, 2nf0, where f0 = 50 or 60 Hz and n = 1, 2, 3, … Intermodulation
frequencies may also appear such as R × N ± 2nf0.

Noise control is the responsibility of the motor manufacturers who have the expertise, but the prediction
of motor noise still remains problematic. Electromagnetic noise is most often observed at low shaft
speed since the aerodynamic noise level decreases much faster with speed than the motor noise.

Use of variable speed control may lead to a high motor noise level when the frequency of the
inverter is in the audible frequency range, i.e., below approximately 10 kHz.

Mechanical noise is similar to electromagnetic noise and may be significant at low speed, unlike
aerodynamic noise. It has the following origins:

-vibrations: vibrations are due to impeller imbalance or misalignment that occurs at harmonics of
shaft frequency. The vibrations are transmitted via the struts to the fan casing, which may radiate
noise as an efficient sound radiator if its dimensions are larger than the acoustic wavelength.

-bearing noise: there are two types of bearings: plain or rolling. Plain bearings are reasonably
quiet if they are lubricated correctly. Rolling element bearings may be noisy if the balls or rollers
are worn.

-coupling: a coupling is not usually a dominant source of noise, but it may transmit vibrations to
adjacent parts. To reduce the amplitude of these vibrations, flexible couplings may be used.

5
Fan Noise

-drives: a belt drive is not an important source of direct noise, but it may transmit vibrations to flat
structures, such as the belt casing, which radiates noise. A casing with wire mesh is less noisy than
a solid cover.

1.2.1.2 Aerodynamic noise

Noise mechanisms

The mechanisms of noise generation in rotating machinery have been extensively studied in the
two following disciplines: aeronautic applications (propellers, fans and compressors in aircraft
engines, helicopter rotors, etc.) and marine propellers. The main mechanisms studied in these two
domains also apply to fan noise. However, due to the diversity in geometry and type of fans, and
due to the impeller tip speed, which is usually much lower on air moving devices than on
aeronautical machinery, hierarchy of the noise sources is different. Deficiencies still remain in the
diagnosis, prediction, and noise control of industrial fans.

Following Lighthill's work on sound generated aerodynamically (1952), Curle (1955) and Ffowcs
Williams and Hawkings (1969) developed theoretical models, which are the basis for the research
done on sound prediction of rotating machinery since Lighthill’s work. Table 1 summarizes the main
noise generation mechanisms of rotating machines (Neise 1988, 767-776). This table is referred
to in the following sections.

The frequency spectrum of fan noise is essentially made up of high amplitude peaks at the blade
passage frequency, BPF = B × N (where B is the number of blades and N is the shaft speed in
cycles per second), and its harmonics, along with a broadband spectrum. Figure 6 shows an
example of sound pressure spectrum between 0 and 1600 Hz, measured at the free inlet of a 400
mm diameter tubeaxial fan with 9 blades and running at 2900 RPM. Discrete tones at BPF and its
first two harmonics are clearly seen on the spectrum. Other peaks also emerge from the
broadband spectrum, but their origin is not clearly understood.

90
BPF
80
2 BPF
70 3 BPF
Lp (dB)

60

50

40

f (Hz)
30
200

400

600

800
0

1000

1200

1400

1600

Figure 6 - Narrow Band Sound Pressure Spectrum of a Tubeaxial Fan

6
General Remarks on Fan Noise

Table 1 - Mechanisms of Aerodynamic Noise Generation of Fans

FAN NOISE
Discrete (D) + Broadband (B)

Monopole Dipole Quadrupole


Blade thickness noise Blade loading noise Turbulence noise
(D) (D + B) (D + B)

Steady loading Unsteady loading


(D) (D + B)

Periodic unsteady Ingestion of


Tip clearance noise
loading turbulence
(B)
(D) (B)

Vortex shedding
Trailing edge noise Rotating stall
noise
(B) (narrow band)
(B)

According to Lighthill's acoustic analogy, one can break up aerodynamic noise sources of rotating
machines into monopole, dipole, and quadrupole sources (1952, 564-587). Table 1 and the
following description of fan noise mechanisms distinguish between discrete tones and broadband
noise origins.

Monopole noise, or thickness noise, is due to the fluctuations of flow resulting from the passage
of the blades in the airflow. This mechanism is directly connected to the displaced air volume, and
thus, with the blade thickness; it also radiates only discrete tones. It can be shown that monopole
sound only becomes significant when the blade tip Mach number, Mt, is greater than about 0.5. For
most fans, this noise is negligible.

Quadrupole sound, which radiates discrete and broadband noise, comes from the fluctuating
shear stress in the flow surrounding the blades. As it becomes important for Mt ≥ 0.8, it is once
again negligible for fans.

Dipole noise is due to the aerodynamic forces exerted by the flow on the rotating blades and
stationary components of the fan. This is the only important noise source for fans. However, it
covers many aspects and is divided into the following categories, referring to Table 1.

7
Fan Noise

Steady loading

The simpler dipolar source of rotating machines is steady loading noise, also called Gutin sound.
This noise, due to the steady aerodynamic lift on the blades, results in discrete tones at BPF and
its harmonics. It can be shown that the noise is negligible compared to unsteady loading noise for
fans of moderate tip speed. Like monopole sound, noise only becomes significant for tip Mach
numbers higher than about 0.5.

Unsteady loading

As seen in Table 1, unsteady loading noise has very diverse origins. These origins lead to both
discrete and broadband noise generation, defined in the following six categories:

Periodic unsteady loading noise (discrete tones)

Periodic unsteady loading noise constitutes the main mechanism of fan periodic noise
generation. This phenomenon occurs when the mean flow velocity at the impeller inlet is not
uniform, especially in the circumferential direction. By referring to Figure 7 (see also definitions
in Annex A), this space non-uniformity of the absolute inlet velocity, Va, induces a variation of
incidence of the relative velocity, Vr, on the blade during the rotation, and therefore leads to a
periodic variation of the resultant mean lift, F, on the blade. It can be shown that this fluctuating
force, F, generates discrete tones at BPF = B × N and its harmonics, given multiples of the
rotational frequency, N, and number of blades, B.

This inconsistency in flow at the impeller inlet can have many origins: wakes of upstream
obstacles such as motor struts or inlet guide vanes; flow non-homogeneity due to a bend or an
asymmetry of the inlet ducting; swirling flow; etc.

Figure 7
Velocity Triangle at the Fan Inlet
F
Va: absolute flow velocity
Ve: blade velocity at radius considered
Vr: relative flow velocity

Va
Vr
Ve

Stationary obstacles at the impeller outlet, such as outlet guide vanes, struts, or the volute cut-
off of centrifugal fans may also be at the origin of discrete tones from a similar mechanism: the
rotating wake of each blade interacts with the fixed downstream obstacle (Figure 8).

8
General Remarks on Fan Noise

Flow velocity in
Figure 8 the blade wake Outlet guide vane
Velocity Triangle at the Fan Outlet

Impeller

Direction of rotation

This noise, which is due to the rotor-stator interaction, has been extensively studied, especially
in aeronautic applications such as axial compressors (Tyler and Sofrin 1962) or helicopter tail
rotors (Fournier 1988). The frequencies of the discrete tones occur at BPF and its harmonics;
i.e., at multiples of the impeller blade number. As shown in Section 1.3.1, the number of outlet
guide vanes does not influence the frequency of the tones, only their amplitude. Tonal noise is
higher when the upstream or downstream obstacles are closer to the impeller. In Section 1.3.1,
we shall give recommendation on the minimum distance that shall be kept between the impeller
and the fixed obstacles to minimize the noise level.

Noise due to ingestion of turbulence (broadband)

This phenomenon is similar to the one described just above. The fluctuations of the turbulent
flow velocity at the fan inlet lead to random fluctuations in the angle of attack, resulting in
random fluctuations of the aerodynamic forces on the rotating blades. In this case, the impeller
radiates a broadband noise, the level of which increases with the inlet turbulence intensity. The
scale of the turbulent velocity structure also influences the shape of the fan sound power
spectrum in a way that depends on the space correlation of the sources over the blades: large
scale turbulent structures with long axial vortices are responsible for narrowband noise
centered on harmonics of BPF, while small scale turbulent structures lead to random noise.

Rotating turbulent wakes of the blades also contribute to broadband noise emission when they
interact with downstream fixed obstacles such as vanes, struts, or the volute cutoff of
centrifugal fans. To reduce this interaction noise, a minimum distance must be kept between
the impeller and the upstream and/or downstream fixed obstacles, and the intensity of
turbulence in the inlet flow has to be as small as possible.

Trailing-edge noise (broadband)

This noise occurs when the turbulent boundary layer on the blade side (usually the suction side)

9
Fan Noise

is convected past the trailing edge. As a result, the turbulent energy of the wall pressure
fluctuations is converted into acoustic energy that radiates to the far-field. The boundary layer
can be attached or separated according to the fan operating point, which will induce a
difference in shape and amplitude of the noise spectrum. However, the sound mechanism
remains basically the same in both cases. Flow separation, which occurs at high load condition,
i.e., at low flow rate on the fan characteristic, generates a spectrum of higher level and lower
frequency than an attached boundary layer.

Trailing-edge noise is the major broadband noise source on low-pressure axial-flow fans when
the inlet turbulence is moderate (Fukano, Kodama and Senoo 1977, 63-74)

Another phenomenon occurs on low speed fans when the boundary layer on the pressure or
suction side is laminar, for example, when the Reynolds number based on the chord length and
relative velocity is much less than 106. The laminar instabilities convected within the boundary
layer may interact with the trailing edge and, due to a feedback mechanism, generate a
narrowband noise at high frequency. These phenomena are discussed in later sections.

Vortex shedding noise (broadband)

This noise source, like the previous mechanism, is located near the blade trailing edge. It is
created by vortex shedding associated with von Karman vortex streets in the blade wakes.
Vortices induce random fluctuations of the blade lift, causing a narrowband noise to be radiated
at a Strouhal number based on blade thickness and mean velocity of about 0.2. Because the
mean relative velocity of axial fans varies along the blade span, vortex shedding noise source
is broadband type rather than narrow-band. Conversely, narrowband noise due to vortex
shedding is more likely in centrifugal fans.

Tip clearance noise (broadband)

For axial fans, the flow at the blade tip is strongly influenced by the tip clearance mechanism
associated with the pressure difference between the suction and pressure sides of the blade.
An interaction between the tip vortex on one blade with the adjacent blades may occur, which
leads to random fluctuations of the blade lift and broadband noise radiation. A reduction of the
tip gap is often an efficient means of noise control, but it is not always the case, especially on
propeller fans, for reasons that are explained later.

Flow separation associated with rotating stall (narrow-band)

This phenomenon, which may occur when the fan is operated at very low flow rates, leads to
emission of narrowband noise or even discrete tones at harmonics of a frequency equal to 0.3
to 0.5 times the blade passage frequency (Neise 1988, 767-776).

10
General Remarks on Fan Noise

1.2.2. Similarity laws and prediction of fan sound levels

1.2.2.1 Fan similarity laws

Introduction

The advantage of using aerodynamic and acoustic fan similarity laws allow the prediction of the
performance of a fan at any speed or impeller diameter from measurements made at one speed
and one diameter. This significantly reduces the amount of testing required. Fan manufacturers
may also use similarity laws to deduce the performance and sound level of large fans from
measurements on scale models.

Fan laws can also be used to assess the influence of fan geometry on the aerodynamic and
acoustic performance rating of different fans of the same type. It is essential to compare the
performance of these fans at the same speed and same diameter, whereas they were tested at
different scales and speed conditions.

Fan laws finally make it possible to simplify the presentation of fan performance curves by
replacing several curves corresponding to various speeds and diameters with one single curve at
reference conditions.

Photo courtesy of Loren Cook Company

Aerodynamic similarity laws

The following equations allow the calculation of the performance of a fan that has impeller
diameter, D2, and speed, N2, from tests performed on a fan of diameter, D1, and speed, N1. This

11
Fan Noise

transposition is valid if the two fans are homothetic or geometrically similar, and if the ratio D1/D2
(or D2/D1) does not exceed approximately 3, to avoid scale effects associated with Reynolds
number. Flow density may also change between conditions 1 and 2.

N2 D23
Qv2 = Qv1 Equation 1
N1 D13

N22 D22 ρ2
Δp2 = Δp1 Equation 2
N12 D12 ρ1

N23 D25 ρ2
P2 = P1 Equation 3
N13 D15 ρ1

η2 = η1 Equation 4

Where Qv, Δp, P and η are volume flow, pressure, power and efficiency of the fan.

The above formulas only apply if the effects of flow compressibility may be neglected, i.e., for fan
pressures below 2500 Pa. For higher pressures, compressibility factors must be introduced into
the equations.

Flow and pressure coefficients are defined from similarity laws, which can be used to plot fan
curves in non-dimensional coordinates.

Acoustic similarity laws

The equation most used to transpose fan sound levels from one condition to another dates back
more than fifty years (Madison 1949). It does not satisfy dimensional analysis but results from
measurements made on many fans of different types and is written as:

Qv2 Δp
Lw2 = Lw1 + 10 lg + 20 lg 2 Equation 5
Qv1 Δp1

In this expression, sound power level, Lw, stands for either the overall level, or the level in octave
or one-third octave bands, provided that fan speed does not change too much between conditions
1 and 2 (see below).

Equation 5 is valid only if aerodynamic similarity laws are fulfilled in the transposition from 1 to 2.
In this case, volume flow, Qv, and pressure, Δp, can be written as a function of speed, N, and
diameter, D, and the following equation, equivalent to Equation 5, may be written:

12
General Remarks on Fan Noise

N2 D
Lw2 = Lw1 + 50 lg + 70 lg 2 Equation 6
N1 D1

One can stress that Equation 6, which is quite general and therefore does not reproduce the proper
evolution of sound level with speed and diameter for any type of fans, is not in accordance with
the trend given by Lighthill's acoustic analogy. For dipole sources like fans, this analogy gives the
following evolution:

N2 D
Lw2 = Lw1 + 60 lg + 80 lg 2 Equation 7
N1 D1

Expressions more sophisticated than Equation 6 have been proposed to take into account the
diversity in sound level of various types of fans. For instance, BS 848, proposes an expression
such as (1985):

N2 D
Lw2 = Lw1 + 10(6 + a )log + 10(8 + 2a + b )log 2
N1 D1

Where a and b are experimental coefficients obtained from tests on different fans.

If the ratio N2/N1 in Equation 6 is not ≈ 1, sound conversion has to be applied not only in amplitude
but also in frequency to account for the distortion of the spectrum with speed; a high speed fan
radiates higher frequency noise than a low speed fan. The following equation is proposed in AMCA
Publication 301 for applying the transposition from conditions 1 to 2 (1990):

N2 D BW2
Lw2 (f2 ) = Lw1(f1 ) + 40 lg + 70 lg 2 + 10 lg Equation 8
N1 D1 BW1

N2
f2 = f1 Equation 9
N1

Where:
Lw is the sound power level in octave (or one-third octave) band (dB)
f is the center frequency of octave (or one-third octave) band (Hz)
BW is the bandwidth of octave (or one-third octave) band filter centered on f (Hz)

As shown in Figure 9, this noise transposition is overall well checked on axial-flow fans. It can also
be adapted to other types of fans, as it will be seen. However, it can be observed on the spectra
of Figure 9 that some peaks do not match the others.

Equation 9 assumes that sound conversion frequency is proportional to fan speed. In practice,
when the speed changes, it is often noted that some peaks on the noise spectrum have a

13
Fan Noise

frequency that does not vary. This is shown in Figure 9 where the peaks between 2 and 5 kHz are
ascribable to motor noise. The frequencies of these peaks are constant regardless of the speed,
so that the transposition causes a frequency shift that is not relevant in this case.

80
N1 = 700 RPM
N1 = 860 RPM
70 N1 = 1030 RPM
LW (dB)

60

50

40
100 200 400
800 1600 3150 6300
Hz
Note: The measured one-third octave spectra are converted to a constant speed N = 1000 RPM.
Figure 9 - Sound Power Conversion on a 400 mm Axial-Flow Fan

Similarly, peaks or humps at acoustic resonance frequencies of a centrifugal fan's volute casing
are observed very often. The resonance frequencies do not change with fan speed, therefore, the
transposition would be still incorrect at these frequencies. To avoid this problem, methods were
developed to separate the contribution of aerodynamic noise in the fan noise spectra, whose
frequencies vary proportionally with speed, from the contribution of other phenomena, the
frequency of which is not related to rotation. One can refer to the work initiated by Weidemann
(1971) and continued by Reznicek and Mongeau (1995, 265-268).

These authors show that the narrowband sound power spectrum of a fan in a system can be
written as:

W ( f ,φ )
= F 2 (St ,φ )G 2 (He )
D3 Equation 10
( ρU ) 2 2

ρ cU

Where:
W = sound power spectrum at the fan operating point defined by the flow coefficient φ (W)
φ = Qv / (πD2U) flow coefficient
ρ = air density (kg/m3)
c = sound speed (m/s)
D = impeller diameter (m)
U = πDN blade tip speed (m/s)
St = f / (B × N) Strouhal number

14
General Remarks on Fan Noise

B = blade number
N = fan speed (Hz)
He = fD/c Helmholtz number

In Equation 10, F is the source function that depends on Strouhal number, St, and flow
coefficient, φ. F accounts for the aerodynamic noise linked with fan rotation, independent of
phenomena related to the sound propagation in the duct system, such as acoustic or mechanical
resonances, or non-aeroacoustic sources. H is the acoustic system response function that
depends on the non-dimensional frequency, He. This function accounts for noise mechanisms not
linked with fan speed. H should not depend on the fan operating point.

This expansion of the fan noise spectrum into two functions, one related to the aerodynamic noise
and the other to the response of the system, is quite useful for analyzing experimental data or for
transposing fan noise spectra to other speed conditions. For a given system resistance curve, this
method may be implemented by measuring narrow-band sound pressure spectra on the same fan
at about ten different speeds, and by applying a signal analysis technique described by Mongeau,
Thompson, and McLaughlin (1995, 369-389).

1.2.2.2 Sound level prediction

This section presents methods of fan sound level prediction based on experimental data tables and
similarity laws. These rough estimates do not take into account the detailed geometry of the fans
or their installation conditions, but they give an idea of the noise levels that one can expect
according to the type of fan.

The ASHRAE method is based on the following equation that is deduced from Equation 5:

Qv Δpt
Lw (f ) = K w (f ) + 10 lg + 20 lg Equation 11
Qv ref Δpt ref

Where:
LW(f) = fan sound power level in octave band radiated by inlet + outlet + casing (dB)
KW(f) = specific sound power level in octave band (dB)
QV = volume flow (m3/s)
ΔpT = fan total pressure (kPa)
Qv ref = 1 m3/s
ΔpT ref = 1 kPa

15
Fan Noise

Table 2 - Specific Sound Power Levels (KW) (Graham 1992, 293-300)

Octave Band Center Frequency (Hz)

Fan Type Diameter 63 125 250 500 1k 2k 4k 8k BPF


Centrifugal Fans
Backward-curved blades > 0.75 m 85 85 84 79 75 68 64 62 3
< 0.75 m 90 90 88 84 79 73 69 64 3

Radial blades: Δp = 1 - 2.5 kPa > 1m 101 92 88 84 82 77 74 71 7


< 1m 112 104 98 88 87 84 79 83 7

Radial blades: Δp = 2.5 - 5 kPa > 1m 103 99 90 87 83 78 74 71 8


< 1m 113 108 96 93 91 86 82 79 8

Radial blades: Δp = 5 - 15 kPa > 1m 106 103 98 93 91 89 86 83 8


< 1m 116 112 104 99 99 97 94 91 8

Forward-curved blades All 98 98 88 81 81 76 71 66 2

Axial Fans
Vaneaxial
Hub ratio = 0.3 to 0.4 All 94 88 88 93 92 90 83 79 6
Hub ratio = 0.4 to 0.6 All 94 88 91 88 86 81 75 73 6
Hub ratio = 0.6 to 0.8 All 98 97 96 96 94 92 88 85 6

Tubeaxial > 1m 96 91 92 94 92 91 84 82 7
< 1m 93 92 94 98 97 96 88 85 7

Propeller All 93 96 103 101 100 97 91 87 5

Table 2 gives the specific sound power levels Kw in octave bands for the most used fans. For each
fan, column BPF shows the value in dB to add to Kw in the BPF octave band to account for the
peak at this frequency.

Kw levels shown in Table 2 were obtained from a large amount of tests performed on well designed
fans at their peak efficiency (Graham 1992, 293-300). Levels should be higher if the fans do not
run at their maximum efficiency. Furthermore, this prediction does not apply for low-pressure fans,
i.e., for fan total pressure lower than 125 Pa.

Neise compares the specific in-duct sound power levels of ten fans of various types on the whole
performance curve (1989). He shows that backward curved centrifugal fans have the lowest
specific sound level, axial-flow fans the highest specific level, and forward curved centrifugal and

mixed flow fans rate between these two extremes. Ponsonnet also gives the evolution of the
specific overall sound power levels of various types of fans along the performance curve (1972).

Bommes suggests a prediction method based on the acoustic efficiency concept (1992, 598-604):

16
General Remarks on Fan Noise

ηacou = Wt / P = 10-4 Reα Mγ -3 Equation 12

Where:
Wt = overall sound power level (W)
P = fan air power (W)
Re = ρcD/μ Reynolds number
M = πDN/c Mach number
c = sound speed (m/s)
μ = dynamic viscosity (kg m-1s-1)

Values of α and γ in Equation 12 depend on fan type, diameter and speed. Bommes provides
values of α and γ deduced from experiments, which allows estimation of sound power levels in
octave band of various fans of different types.

Prediction methods presented in this section only give rough estimations of the fan noise level with
an uncertainty of about 2 to 5 dB(A). The uncertainty in each octave band may be even higher. To
improve this prediction, it is recommended that a database be developed from tests performed on
various fans using the same measurement method and applying fan similarity laws.

These methods are based on statistical results obtained from tests and not on a real understanding
of the physical phenomena which could lead to an actual prediction. Therefore, they are useful for
noise level estimation, but essentially useless for improvements in fan design.

1.2.3 Determination of fan sound levels using standardized methods

1.2.3.1 Introduction

Standardized methods for the determination of fan noise level have been developed for a long time
in several countries. For example, in France, NF S31-021 was well suited to measure the sound
power level at the free inlet. Countries like USA, UK, or Germany also had their own test methods.
Applying each of these methods to the same fan would not necessarily give the same results,
therefore, it became necessary to set up an international standard for fan noise testing.

In 1992, ISO/TC 117, in charge of industrial fans, decided to draw up a fan noise test standard from
the existing national standards. An effort had to be made for improving and harmonizing these
different methods.

ISO Standard 13347 provides several methods for measuring fan sound power level in one-third
octave bands (or octave bands), taking into account fan installation. According to ISO terminology,
four categories of installations are recognized: A, B, C and D (Figure 10), and for each category,
up to three fan sound levels may be defined:

-inlet level
-outlet level
-fan casing (significant only for type D)

17
Fan Noise

Figure 10
Fan Installation Categories

Type A Type B

Type C Type D

ISO 13347 describes three test methods:

-reverberant room method


-enveloping surface method
-sound intensity method.

Two other methods that apply to fans are mentioned but not detailed in ISO 13347 as they are
described in other ISO standards:

-in-duct method (ISO 5136)


-plastic plenum method (ISO 10302).

1.2.3.2 Reverberant room method

The reverberant room method is based upon the AMCA Standard 300 (1996). Tests are performed
in a reverberant room where the sound pressure field is nearly homogeneous in the whole space
for frequencies between 50 and 10,000 Hz. A procedure for the qualification of the test room is
described in the publication. This room must also be equipped with ancillary equipment to modify
and control the fan operating point.

The test fan is either ducted or not ducted, according to installation category A, B, C or D, and the
sound power level is measured at the fan inlet or outlet. Figure 11 shows an example of the test
installation for measuring the inlet sound power level of a non-ducted fan (type A installation). Fan
pressure is the difference in pressure between the inlet and outlet rooms. Volume flow is to be
measured according to one of the methods specified in ISO 5801, which is the standard for the
determination of fan performance.

18
General Remarks on Fan Noise

Figure 11
Reverberant Room Method Reverberant
room

The measurement of the fan sound power level in one-third octave bands is made using a
reference sound source (RSS), the sound power of which is known from its calibration. Sound
pressure levels are measured with the microphone at several positions in the reverberant room
with only the RSS in operation, then with only the fan in operation. For a given operating point, the
fan sound power level in each frequency band is obtained from:

LWv (f ) = LPv (f ) + LWref (f ) − LPref (f ) Equation 13

Where:
LWv = fan sound power level (dB)
LPv = fan sound pressure level averaged over the microphone positions in the room (dB)
LWref = RSS sound power level (dB)
LPref = RSS sound pressure level averaged over the microphone positions in the room (dB)

1.2.3.3 Enveloping surface method

This method determines the fan sound power levels for the same installation conditions as with the
reverberant room method. Test results obtained from these two methods should be very similar, as
shown by Holste and Neise with measurements on six different fans (Holste and Neise 1992, 1-26).

The method is based on the specifications of ISO 3744, which describes a procedure
to determine the power level of a sound source in a free field over a reflecting plane (1994).

The fan under test is connected to an installation that allows controlling and determining the
aerodynamic performance simultaneously with the acoustic performance. The sound power level
is obtained from the measurement of sound pressure levels at different locations on a surface
enveloping the source. The measurement surface differs according to the original national
standard used. This surface may be a parallelepiped enclosing the fan (or the inlet or outlet), on
which 16 measuring points are defined in accordance with DIN 45 635 Part 38 (1986).

19
Fan Noise

Figure 12
Enveloping Surface Method -

150r
S9
Rectangular Paralellepiped

Do

Dr
h

Dr
SH

r
3D

3Dr Reflecting
floor or wall

y
Figure 13
Enveloping Surface Method - o4
Large Hemisphere
3o
9
o
o5
10
o
-x V o8 x

o
6
o 2o
7

o indicates microphone position o1


V indicatedes positions of fan
-y
z
Ra 10
di o
us o9
of
he o7 o
8
m
isp
o5 he o4
o re 3
o
6 r o1 o
2
-x V x
Reference plane

20
General Remarks on Fan Noise

Figure 14
Enveloping Surface Method -
Small Hemisphere
Z
7 7
8 Reflecting
Plane 8
°
80
°
60

0 1 2 3 0 4 5 6

DH
X Y
33°

9 Reflecting 9
Plane
>1 m

Determining the fan sound power can also be accomplished by using the sound pressure level
measured in 13 positions on a hemisphere surrounding the fan, displayed in Figure 13 (BS 848
Part 2 1985).

Lastly, the measurement surface may be in conformity with NF S 31-021 (1982) and made up of a
small hemisphere centered on the fan inlet (Figure 14). The number of measuring points is only 6
in this case. This method is not suitable to measure the sound power level at the fan outlet.

For these three types of measurement surface, the fan sound power level in one-third octave (or
octave) bands is given by the following equation:

Lw = L p + 10 lg S − K1 − K 2 Equation 14

Where:
Lw = fan sound power level (dB)
LP = fan sound pressure level averaged over the measurement surface (dB)
K1 = background noise level correction (dB)
K2 = environmental correction (dB)

The background noise correction, K1, in each frequency band, is obtained from the difference in
levels measured with and without the fan in operation. If the background noise level is too high,
i.e., if the difference between the fan and background noise levels at each frequency is lower than
6 dB, the test should not be continued until the background level decreases.

21
Fan Noise

The environmental correction, K2, takes into account the fact that the test environment is never a
perfect acoustic free field over a reflecting plane. This method does not require a true anechoic
room. It may be applied, for instance, in an industrial building, provided that the room around the
measurement surface is large enough to satisfy some criteria for K2.

The value of K2 in each one-third octave band is obtained from pressure level measurement with
a reference sound source (RSS) located at a position specified in the standard, according to:

K 2 = L Pref + 10 lg S − LWref Equation 15

Where:
LPref = RSS sound pressure level averaged over the measurement surface (dB)
LWref = RSS sound power level (dB)
S = area of the measurement surface (m2).

According to ISO 3744, the test environment is suitable if –2 ≤ K2 ≤ 2 dB, while in an "ideal" free
field, K2 = 0 (1994). In practice, it is difficult to keep K2 within these limits in the whole frequency
range [50 – 10000 Hz], but, if this criterion is not fulfilled at a few frequencies, it has been shown
that the noise level measured on several fans was only slightly inaccurate (Gray 1994).

Figure 15 shows a comparison of the sound power spectra at the inlet of a centrifugal fan (320 mm
impeller diameter) which was obtained from the pressure levels measured on the three
measurement surfaces described above (Guedel and Freynet 1997a). The discrepancies between
the spectra are maximized at 4 dB, with an average of about 2 dB, which remains acceptable
according to the uncertainties specified in ISO 13347.

90
DIN NF BS
85

80

75
LW (dB)

70
65

60
55
50 100 200 400 800 1600 3150 6300
f (Hz)
Figure 15 - Compairson of Sound Power Spectra Measured
According to DIN, BS, and NF Methods

22
General Remarks on Fan Noise

1.2.3.4 Sound intensity method

This method measures the same quantities as the methods described above. For the same fan
and the same installation conditions, these three methods should give very similar results. The
advantage of this last method compared to the two others is that it does not require a specific test
environment and that it is less sensitive to background noise. Nevertheless, it requires a sound
intensity measurement instrumentation and a more qualified person for conducting the tests.

The sound power level in each frequency band is deduced from the sound intensity distribution on
a measurement surface around the fan. This surface may be either a hemisphere or a
parallelepiped, and the intensity levels may be measured at discrete points or by scanning
according to ISO 9614-1 (1993) or ISO 9614-2 (1993).

1.2.3.5 In-duct method

Unlike the three previous methods, this method measures fan in-duct sound power levels for type
B, C and D installations. The method is detailed in ISO 5136 (1999). The fan sound power level is
deduced from the sound pressure levels measured at three positions in a duct cross-section. The
duct has to be fitted with an anechoic termination to avoid the stationary waves due to the
interference between the direct wave radiated by the fan and the reflected wave at the duct end
(Figure 16).

This standard recognizes that acoustic and fan performance measurements are to be made at the
same time; therefore, the duct has to be equipped with test arrangements to control and measure
the operating conditions of the fan (e.g., an orifice plate; see Figure 16). This also requires that the
common part, as defined in ISO 5801, is introduced at the fan inlet and/or outlet.

Figure 16
In-Duct Method fan in-duct sound pressure anechoic
measurement termination

v x
pressure
taps flow control

The in-duct microphone must be protected from the flow fluctuations by a sampling tube made
from a metal tube with a longitudinal slit (Figure 17). This device enables the measurement of fan
acoustic pressure fluctuations while decreasing the amplitude of both the self-induced wind noise
on the microphone and the turbulent pressure fluctuations of the airflow within the duct.

23
Fan Noise

Figure 17
Sampling Tube for a 1/2 Inch Microphone
Microphone with Slit
Porous material Slit protection grid 1

12,7
15

1
400

When the mean flow velocity within the duct is lower than 15 m/s, a nose cone or a foam ball may
be used instead of the sampling tube.

The in-duct fan sound power level in each one-third octave frequency band is obtained from:

Lw = Lp + 10lg S + C Equation 16

Where:
LW = fan sound power level (dB)
Lp = fan sound pressure level averaged in the duct cross-section (dB)
S = area of the duct cross-section (m2)
C = correction that depends on the type of microphone shield, duct diameter, flow velocity and
frequency (dB).

1.2.3.6 Test plenum method

This method, which is described in ISO 10302 (1996), is well suited for the determination of the
sound power level of small fans, such as air-moving devices used for cooling electronic equipment
and similar applications. The test fan is connected to a cubic test plenum chamber of 1 meter on
a side, constructed with a frame covered by an airtight, acoustically transparent Mylar film, a
mounting panel, and an adjustable exit port assembly, as shown in Figure 18. The adjustment of
the position of the slider on the exit port assembly enables the fan operating point to be varied,
whereas the measurement of the static pressure within the plenum for a given slider opening
accounts for this operating point.

This method applies to fans for which the maximum airflow is 1 m3/s, and the maximum static
pressure is 750 Pa. The fan outlet diameter is also limited to 500 mm.

24
Parameters of Influence, Modeling, and Control of Fan Noise

Figure 18
Test Plenum [34]

Adjustable exit port assembly


Mylar® sheets Slider opening
all areas, including
bottom, except mounting
panel and exit port Slider

0.3 M

0.6 M
0.9 M

0.6 M
0.3 M
Fane

0.1 M
Typ places
Piezometer pressure 1.2 M
ring behind panel 1.2 M Gusset and vibration
Mounting panel assy isolation
Retainer
Fan

1.3. Parameters of Influence, Modeling and Control of Fan Noise

1.3.1. Axial-flow fans

1.3.1.1 Introduction

There are three types of axial-flow fans:

-vaneaxial
-tubeaxial
-propeller

This section describes the main geometric and performance characteristics, parameters that have
some influence on noise, sound prediction models, and noise control methods of axial-flow fans.
Circulating fans, such as ceiling or table fans, are not covered since their pressure rise is nearly
zero and it is almost impossible to dissociate the inlet from the outlet side.

1.3.1.2 Fan design and air performance

Vaneaxial fans have an impeller with a hub/tip ratio that is usually 50% or more, and a set of
downstream guide vanes (or more seldom, upstream guide vanes), the rotor and the stator being

25
Fan Noise

surrounded by a casing (Figure 19). The blades of the impeller are of airfoil type and twisted, i.e.,
their pitch angle varies along the span. These fans are very aerodynamically efficient; the total
efficiency may reach 90% for well-designed vaneaxial fans (Daly 1992). Fan pressure is higher
than on the other axial fans, as the outlet guide vanes increase the pressure produced by the
impeller. It may reach several thousands of Pa.

Figure 19 impeller outlet guide vanes


Vaneaxial Fan

Tubeaxial fans have blades with shapes similar to those of vaneaxial fans, but they are not
equipped with guide vanes (Figure 20). Blades are usually of airfoil type, but they may also be of
constant thickness. The hub ratio is usually lower than 50%. Total efficiency of these fans may
reach 80%, and fan pressure does not exceed 1000 Pa.

Figure 20
Tubeaxial Fan

The number of blades is between 3 and 10. The fan shroud may be quite short (similar to the axial
size of the impeller). This fan may be installed according to any type of installation; A, B, C or D.

Tubeaxial and vaneaxial fans should operate close to their best efficiency point for aerodynamic
and acoustic reasons; i.e., at 60 to 80% of their maximum volume flow.

Propeller fans have 2 to 6 blades. The blades, most often of constant thickness, are usually wide
at the tip. Their shape varies from simple rectangular sheet to the larger swept shape as shown in
Figure 21. This last blade geometry is mostly used on small diameter fans (D ≤ 500 mm). The width
of the shroud surrounding the impeller is a fraction of the blade axial tip chord length. The width
and the shape of this ring, as well as the axial location of the impeller in the ring, have a significant

26
Parameters of Influence, Modeling, and Control of Fan Noise

influence on the fan performance.

The maximum static pressure of propeller fans is 100 to 200 Pa, which is sufficient for applications
such as air conditioning units and cooling towers, where these fans are widely used. They operate
satisfactorily in the flow range between 40 and 100% of the maximum flow, and their total efficiency
is not more than 60%.

Figure 21
Propeller Fan

1.3.1.3 Description and prediction of sound generation mechanisms

Mechanisms of fan sound generation have been briefly described in Section 1.2.1. Those
concerning axial-flow fans, as well as the parameters that have a significant influence on noise,
are now detailed. Guedel provides the description of the physical phenomena (1989, 1993).

In Section 1.2.1, the role of the periodic and random blade forces on the radiated noise has been
highlighted. A complete prediction of these forces, especially the random forces associated with
the turbulent inflow, and an accurate calculation of the radiated noise level, exceed the possibilities
of current computers, but simplified models have been developed, some of which are presented
here.

Tonal noise

For all axial-flow fans, tone noise at BPF = B × N and its harmonics comes from the periodic
fluctuating forces on the blades induced by the non-homogeneous, steady flow velocity at the fan
inlet. On vaneaxial fans, a second important source is due to the interaction of the rotating blade
wakes with the outlet guide vanes.

A theoretical prediction of the tone amplitude has been established by Lowson (1970, 371-385)
from earlier work by Ffowcs Williams and Hawkings (1969). The formula below gives an
expression for the complex sound pressure of the BPF tone radiated by an axial fan in the far field:

27
Fan Noise

iB 2 Ω λ =∞ ⎛ B−λ⎞
p(BΩ) = ∑ (−i )B −λ ⎜⎝ Fx (λ )cosψ − Ft (λ ) BM ⎟⎠ JB −λ (BM sinψ )
2π c0 r1 λ =−∞
Equation 17

Where:
B = number of blades
Ω = 2πN
N = shaft rotation frequency (Hz)
c0 = sound speed (m/s)
r1 = distance between the fan hub and the observer (m)
λ = harmonic rank of the periodic load on the blade (λ = 1, 2, 3, …)
Fx = axial component of the periodic resultant lift on one blade
Ft = tangential component of the periodic resultant lift on one blade
ψ = angle between the impeller axis and the direction of observation (Figure 22)
M = tip rotation Mach number
JB-λ = Bessel function of order B-λ

Figure 22 z
Coordinate System

ωR
source

o y

r
fan axis
x r1

fixed observer

The main difficulty in predicting fan sound pressure levels from Equation 17 is estimating the
amplitude of the periodic fluctuating forces at the harmonics of the shaft rotation frequency, λN.
Much work was undertaken to estimate these force fluctuations from the flow velocity pattern in the

28
Parameters of Influence, Modeling, and Control of Fan Noise

impeller inlet section. One may refer to the earlier work of Kemp and Sears and of Horlock (1955;
1968), the objective being to assess the complex transfer function between the inlet mean velocity
pattern and the aerodynamic forces on the blades (see Section 2.2.3).

Lowson proposes a very simple model to estimate these unsteady forces from the blade steady
load (1969). He assumes that the amplitude of the λth harmonic of the periodic load decreases
exponentially when λ increases: FX(λ) = FX0λ-h where FX0 is the axial steady force on the blade,
and h is an arbitrary constant function of the blade geometry. This unsteady load modeling can only
give a rough prediction of the amplitude of the BPF peaks. Comparisons between prediction with
this model and measurement are shown by Bridelance (1982); Marteel, Desmet, and Wullens
(1992); and Ameziane, Pauzin, and Guilhot (1992).

These fluctuating forces may also be estimated from the measurement of wall pressure
fluctuations with transducers rotating with the blades; however, the implementation of this
technique still remains tricky and expensive. Prediction of the unsteady load by CFD techniques is
being used more and more with some success.

On vaneaxial fans, the interaction of the rotor blade wakes with the outlet guide vanes is also an
important source of periodic noise. The stator radiates tones at BPF (and its harmonics) and not
at a frequency multiple of the stator vane number.

Lowson shows that the expression of the far field sound pressure emitted by the stator is quite
similar to Equation 17, relating to the rotor (1970). The amplitude of the peak radiated by the stator
at BPF is:

iBV Ω λ =∞ ⎛ B − kV ⎞
p(BΩ) = ∑
2π c0 r1 λ =−∞
( −i )B −kV ⎜ Fx (B )cosψ − Ft (B )
⎝ BM ⎟ JB - kV (BM sinψ )

Equation 18

Where:

Fx = axial component of the periodic resultant lift on one stator vane


Ft = tangential component of the periodic resultant lift on one stator vane
V = number of stator vanes
k = 0, ±1, ±2, ±3, .....

Other symbols are identical to those used in Equation 17, angle ψ and distance r1 being defined in
this case from the stator inlet section. Unlike in Equation 17, only harmonic B of the unsteady load
on the stator vanes contributes to the radiation of the stator periodic noise.

Similar to the rotor, the main difficulty in the stator noise prediction is the assessment of the
unsteady forces on the vanes due to the rotating wakes of the blades. These forces may be
measured using wall-pressure probes on the vanes (it is easier to measure fluctuating pressure
signals on fixed rather than on rotating airfoil), or they may be calculated from the blade wake
obtained either by testing or by CFD calculation. This last approach is presented by Fournier in the

29
Fan Noise

case of a helicopter tail rotor where the noise phenomena are similar to those occurring on fans
1988).

In the same reference, another mechanism is addressed: the "potential interaction", which is due
to the presence of fixed obstacles, such as struts or vanes, in close proximity to the rotor outlet.
The disturbance of the mean flow due to these obstacles propagates upstream to the rotor, which
radiates tones at harmonics of BPF because of the periodic load on the blades associated with this
flow disturbance.

It can be shown that the "potential interaction" is a more efficient acoustic radiator than the "viscous
interaction," which is the interaction of fixed obstacles with upstream rotating wakes, described
above (Kaji and Okazaki 1970). The amplitude of the peaks associated with the potential
interaction decreases very quickly when the distance from the rotor to the downstream obstacles
increases, while the tones due to the viscous interaction very gradually decrease as this distance
increases. It can be considered that when the axial clearance between the rotor and the stator is
larger than 0.3 C, where C is the blade chord length, the viscous interaction provides a greater
contribution to the tonal noise than the potential interaction. This remark emphasizes the
importance of the rotor-stator distance on the noise level of a vaneaxial fan, and it also explains
why a distance of at least 0.3 C between the two rows is required for noise control.

Photo courtesy of Greenheck Fan Corporation


Broadband noise

The broadband noise mechanisms on tubeaxial, vaneaxial, and propeller fans are similar, even if
the importance of the different sources may differ from one fan to another. When downstream
guide vanes are present, an additional source may occur, which will be explained later.

The four main broadband noise sources of axial-flow fans are:

-interaction of blades or vanes with inflow turbulence


-trailing-edge noise due to the convection of attached or separated turbulent boundary layer past
the blade trailing edge
-vortex shedding noise associated with von Karman vortex streets in the blade wakes
-noise generated by leakage flows and tip vortices.

30
Parameters of Influence, Modeling, and Control of Fan Noise

The second and third sources make up blade self noise, while the first and the fourth sources form
the interaction noise.

As seen in Section 1.2.1, these mechanisms are all related to random fluctuating forces on the
blades, which constitute the main input data to noise prediction models. These data are not easy
to obtain, therefore, simplified models have been developed to try to predict broadband noise
levels emitted by axial fans.

One of these models is due to Sharland (1964), who showed that the overall sound power due to
the random load fluctuations on one blade is given by:

2
ρ0 ⎡ ∂CL ⎤
P= ∫env CW SC < ⎢⎣ ∂t ⎥⎦ > dr
4
Equation 19
48π c03

Where:
P = overall sound power of one blade (watts)
c0 = speed of sound (m/s)
ρ0 = flow density (kg/m3)
C = blade chord length (m)
W = relative mean flow velocity (m/s)
Sc = correlation surface of the pressure fluctuations on the blade (m2)
CL = unsteady lift coefficient of the blade
r = spanwise coordinate
env = span
< > = time average

Equation 19, which relates the overall sound power radiated by the blade to the mean squared
value of the time derivative of the lift fluctuation, is quite general since it applies to the four
mechanisms mentioned above. To predict the sound levels due to one particular mechanism,
appropriate data have to be entered into a modified equation. Sharland’s formula, which predicts
the overall level of the trailing-edge noise of a low-pressure tubeaxial fan, is modified to show the
role of the blade wake thickness on the fan sound power (Fukano, Kodama, Senoo 1977).

πρ0 B
1200c03 ∫
P= D(r )W 6 (r )dr Equation 20

Where:
P = overall fan sound power (W)
B = number of blades
c0 = speed of sound (m/s)
ρ0 = flow density (kg/m3)
D = wake thickness close to the blade trailing edge (m)
W = relative mean flow velocity outside the wake (m/s)

31
Fan Noise

In Equation 20, the integration on r is made along the blade span. D is the sum of the blade
thickness and the boundary layer thickness on the suction and pressure sides at the trailing edge.
For attached turbulent boundary layers, D can be written at any radius r:

D = Dt + 0.09C.Rec −0.2 Equation 21

Where:
Dt = blade thickness at the trailing edge (m)
C = chord length at r (m)
Rec = Reynolds number based on chord C and relative velocity W

D may also be obtained from mean velocity measurement with a hot wire in the wake close to the
blade trailing edge (the hot wire signal being synchronized with the blade rotation) or from CFD
calculation.

Equation 20 shows that to reduce fan trailing-edge noise, which, according to Fukano, is the major
broadband noise source of low-pressure axial fans when the inlet flow is weakly turbulent, the
wake thickness has to be as small as possible and flow separation on the blade must be avoided.

Carolus deduces from Sharland's formula another expression for the trailing-edge noise prediction
(1992, 802-815):

P ∝ B ∫ C(r )W 6 (r )Rec −0.4 (r )dr Equation 22

Using the same symbols as Equation 21. Equation 22 is similar to Equation 20 when D is replaced
by its expression given in Equation 21.

Prediction Formulae Equations 19, 20, 21, and 22 are easy to implement, and they provide useful
information for noise control (see Section 1.3.1.4). Their drawback is that they only give an overall
sound level prediction in the frequency range where the specified noise mechanism is
predominant. A prediction of the sound spectrum would be more valuable. Some authors proposed
extensions of Sharland's formula to account for a frequency dependence.

Besides these simplified models, more elaborate approaches have been implemented to predict
noise levels associated with specific mechanisms. For instance, models for the prediction of blade
trailing-edge noise were proposed (Howe 1978, 437-465; Amiet 1976, 387-393). Without going
into the details of these models, it should be noticed that the input parameters entering the
prediction are the power spectral density and the spanwise coherence scale of the wall-pressure
fluctuations on the blade sides close to the trailing edge.

The determination of the aerodynamic pressure fluctuations can be obtained from experiment with
flush-mounted pressure transducers mounted on rotating blades. This has been done, for
instance, by Fuest and Carolus, who measured the wall-pressure spectra on the blades of two
axial fans (one with conventional and the other with swept blades) with three miniature
piezoresistive pressure transducers at twelve discrete measuring points on the suction and

32
Parameters of Influence, Modeling, and Control of Fan Noise

pressure sides (1995, 97-100). They obtain useful information relating the blade pressure
fluctuations and the far-field broadband acoustic pressure.

This kind of experiment is still expensive and difficult, thus, the determination of the pressure
fluctuations may also be done on fixed airfoil in wind tunnel, as it is much easier to accomplish.
The data are then extrapolated to rotating blades with some assumptions. The use of CFD
calculations to obtain the blade pressure fluctuations on fixed or rotating airfoils does not provide
reliable results yet.

Amiet and Paterson have investigated broadband noise due to the interaction of blades or vanes
with inlet turbulence (1979). Amiet's model uses the power spectral density and the spanwise
coherence scale of the upstream turbulent velocity field as input data. The model includes an
aeroacoustic transfer function between the inlet fluctuating velocity and the blade lift fluctuation,
which determines the radiation efficiency of the oncoming disturbances. The shape of the sound
spectrum depends on the inflow turbulent scales: when these scales are smaller than the distance
between two adjacent blades, the noise spectrum is of broadband type since the noise radiated by
each blade is uncorrelated from one blade to the other. Conversely, when the turbulent eddies are
larger than the blade spacing, the sources are more or less correlated and the spectrum shows
humps centered on BPF and its harmonics. This mechanism is further detailed in Section 2,
relating to acoustic installation effect.

A direct relation between the aerodynamic pressure fluctuations on the blades and the noise
radiated in the far field has been emphasized. These pressure fluctuations cannot be obtained with
enough accuracy yet from CFD computation. Therefore, they have to be experimentally
determined either from direct or indirect measurements as mentioned above. With the recent
progress made in turbulent flow simulation and Computational Aero-Acoustics (CAA), fan
broadband noise prediction should be effective in a not too distant future.

Photo courtesy of Hartzell Fan, Inc.

33
Fan Noise

1.3.1.4. Noise reduction methods

Tonal noise

Tonal noise reduction, which is particularly relevant for high-speed fans, requires methods derived
from the description of the mechanisms in the previous section.

Since discrete noise generation is due to an interaction between the inlet, non-uniform flow field
and the impeller, it is advisable to keep the inlet flow field as homogeneous as possible by keeping
a distance of at least 3 to 4 duct diameters between the fan and inlet obstacles such as struts,
bends, dampers, and quick duct expansion or contraction. Avoiding inlet guide vanes and large
obstructions just before the impeller is strongly recommended.

In the same way, fixed obstacles near the fan outlet, which radiate noise when the rotating blade
wakes act on them, should be moved away from the impeller. A practical guideline is to keep a
distance of at least one rotor blade chord between the impeller and downstream obstacles such
as outlet guide vanes (OGV) or motor struts. This minimum distance may be a good compromise
to preserve reasonably good performance and low-noise levels on vaneaxial fans.

The number of vanes or struts is also an important parameter for noise control, especially when
the distance between the impeller and the downstream obstacles is not large enough. To minimize
the discrete tone levels, the number of rotor blades and stator vanes must satisfy the following
equation:

k m c0
nBN < Equation 23
πD

Where:
n = noise harmonic number (n = 1 corresponds to BPF, n = 2 to 2 BPF, ...)
B = number of rotor blades
N = fan speed (Hz)
c0 = speed of sound (m/s)
D = impeller diameter (m)
km is a non dimensional term that depends on m = |nB - kV|, where V is the number of stator vanes
and k = 0, ± 1, ± 2, ...

- km = 0 for m = 0
- km = 1.84 for m = 1
- km = 3.05 for m = 2
- km = 4 for m > 2

If B = V, the term on the right of Equation 23 is null, so that this inequality is never satisfied, and
the amplitude of the peaks at harmonics of BPF is maximum.

The use of unevenly spaced blades on the circumference is another means of tonal noise control.

34
Parameters of Influence, Modeling, and Control of Fan Noise

This method reduces the amplitude of the tones at BPF and its harmonics, but increases the
amplitude of other harmonics of the rotation frequency N, so that the overall sound power level of
the fan remains unchanged. The replacement of a big peak at BPF by smaller peaks at harmonics
of N might nevertheless be interesting in terms of sound quality.

An increase in the number of blades is beneficial for tonal noise reduction at a constant operating
point, but this increase may be detrimental to the broadband noise level (Bridelance 1982;
Ameziane 1992). This is shown in the next section.

Other means, such as the use of swept blades or highly loaded blades, may be efficient to reduce
rotational noise, as well as random noise as shown in the following section.

Broadband noise

The main methods of fan broadband noise control are now described. Some of them result from
the description of the noise mechanisms presented above. These methods may also apply in some
cases for tonal noise control.

- Axial distance between the impeller and the inlet or outlet fixed obstacles: this method has
already been mentioned for tonal noise, and it also applies for broadband noise, since a reduction
of the inflow turbulence associated with wakes of obstacles or flow disturbances is beneficial for
noise control. We shall develop this point in Section 2.

Similarly, a minimum distance must be kept between the rotor and outlet guide vanes or struts to
minimize the noise due to the impingement of the rotor wake turbulence on these downstream
obstacles. As suggested for tonal noise, a minimum distance of one blade chord length may be
adopted.

- Flow separation: flow separation occurs on the blade suction side at low flow rate on the fan
curve, due to a high angle of attack. In this portion of the fan characteristic, the noise level
considerably increases so that it is advisable to make the fan running out of this area, i.e., closer
to its best efficiency point.

On propeller fans, the radial component of the mean flow velocity may be important, especially at
the impeller outlet in the low flow range. In this case, a strong separation, or even a reversed flow,
may occur in the hub area (Figure 23), which leads to an increase in noise level (Guedel, Perrin,
Freynet 1997).

A modification of the shape of the impeller (or of its shroud) to make the flow field more axial in the
operating range is favorable for noise control.

35
Fan Noise

Figure 23
Flow Pattern in a
Propeller Fan

large flow low flow

If the blade camber angle is too high, a significant increase of the noise level may occur due to
flow separation on the suction side. A highly cambered blade is nevertheless recommended for fan
noise control because an increase of the mean load is beneficial for fan performance and not
detrimental to noise level if flow separation is not reached (Guedel 1993).

- Blade geometry: the blade shape has a very significant role on the fan noise level, in a way that
is not well understood or quantified yet. It is only possible to give trends that are supported by
common sense, experience, or existing prediction models.

A first means to reduce noise at a given operating point is to run at the lowest possible shaft speed.
This may be achieved by choosing blade geometry leading to a high performance curve, in such
a way that for a given operating point, this fan runs at a lower speed than that of other fans of lower
performance. The use of highly loaded blades is beneficial for this reason (Bridelance 1982).

An increase in the tip blade camber is worthwhile too, provided flow separation on the blade sides
is avoided. Tip camber angles may reach values of up to 20° to 40° according to the blade profile
(Guedel 1993; Guedel, Perrin and Freynet 1997). On low-pressure axial fans, it appears preferable
to use blades strongly loaded near the tip rather than blades with a free vortex design, i.e., with a
constant load along the span (Carolus 1992, 809-815).

The influence of parameters such as blade number, chord length, blade sweep, machining of the
trailing edge on noise level, etc., has been studied by different scientists (Guedel 1993). The
following summarizes the main results of this work.

A reduction in the number of blades along with an increase in the tip chord length is beneficial to
noise. This result is in agreement with experiments and with Equations 20, 21, and 22. It can be
shown that the overall sound power varies as BC0.8 (where B is the number of blades and C the
tip chord length) if Equation 20 and Equation 21 are considered, or BC0.6 if Equation 22 is taken.
At constant solidity σ, where σ = BC / (πD) with D fan diameter, i.e., at constant fan work, the sound
power level increases more quickly with the number of blades than with the chord length.

36
Parameters of Influence, Modeling, and Control of Fan Noise

This argument is strengthened if the overall A-weighted sound power level (OASPL) is considered,
since, according to experimental results of Fukano, a fan with a low number of blades of large
chord length radiates less noise in the high frequency range (f > 1 kHz), which is very detrimental
to the OASPL in dB(A), than a fan with a larger amount of blades of narrow chord. This frequency
shift towards the lower frequencies results from raising blade width more than from reducing the
number of blades (Fukano, Kodama, Takamatsu 1977, 75-88)

It can be emphasized that a low blade number is beneficial to broadband noise, but detrimental to
rotational noise as seen above. Therefore, the choice of the blade number should depend on the
respective contribution of the tonal and broadband noise to the overall fan level.

The use of forward swept blades (Figure 24) is an efficient means of tonal noise control because
they cause a phase shift of the sound sources distributed along the blade span. This configuration
is also efficient for reducing broadband noise level, since the sweep induces a reduction of the
random fluctuating load on the blades (Fuest, Carolus 1995, 97-100; Guedel, Yazigi 1992, 167-
178). The actual mechanism associated with blade sweep on broadband noise control still needs
to be completely clarified.

a
Figure 24
Unswept (a) and Swept (b) Blades

To reduce the contribution of trailing-edge noise and vortex shedding noise, solutions such as a
beveled trailing edge (Figure 25) or a serrated trailing edge (Figure 26) were proposed. The effect
of a serrated trailing edge is to reduce the radiation efficiency of the trailing edge when the incident
wall-pressure fluctuations are convected in front of it. Beveling the trailing edge is intended to avoid
the formation of vortex shedding when the boundary thickness on both sides of the blade is thinner
than the blunt trailing edge. Both solutions proved to be very efficient on fixed airfoils in a wind-
tunnel, but not so conclusive on actual machines, such as fans or wind turbines (Burgain 1998).

37
Fan Noise

Figure 25
50°
Beveled Blade Trailing Edge

25°

2h
Figure 26
Serrated Blade Trailing Edge e

- A high frequency narrowband noise may occur when the boundary layer on the suction side
is laminar. This is due to a feed back loop between instability waves in the laminar boundary layer
that are convected past the trailing edge and the acoustic waves radiated by the trailing edge. This
noise source may be avoided by making the blade boundary layer turbulent, i.e., by placing a
tripping device (adhesive tape) on the blade side on which this phenomenon occurs (Bridelance
1986, 141-146). Other means, such as increasing the turbulence of the flow field at the impeller
inlet by a grid may be efficient, provided that the turbulence rise does not lead to an increase in
broadband noise due to the interaction of the turbulent inflow with the blades (Burgain 1998).

- Tip clearance reduction: a reduction of the radial gap between the blade tip and the fan casing
is beneficial to the aerodynamic and acoustic performance on tubeaxial and vaneaxial fans. This
gap reduction leads to a suppression of the tip vortex, which is detrimental to air performance as
well as to tonal and broadband noise. Highly efficient axial fans may have tip clearance of 0.1 to
0.2 percent of the impeller diameter. Nevertheless, when the gap is small, the impeller and the
casing have to be as concentric as possible since an uneven tip clearance, due to a bad
concentricity or an oval casing, may induce a higher noise level than that of the same fan with a
higher tip clearance (Fukano, Takamatsu, Kodama 1986, 291-308). Furthermore, the influence of
the tip clearance depends on the blade chord length: when the tip chord length is wide, the
clearance has less influence (Hunnaball 1992, 475-482).

On propeller fans, experiment shows that a reduction of the tip clearance does not improve the
aerodynamic and acoustic performance (Guedel 1995a). The flowfield in these fans is different
from the one encountered in tubeaxial and vaneaxial fans because of the strong radial component

38
Parameters of Influence, Modeling, and Control of Fan Noise

of the flow velocity at the fan inlet or outlet according to the operating point (see Figure 23). This
flow pattern, and the fact that the fan shroud is noticeably narrower than the axial chord length,
may explain the small influence of the tip clearance on propeller fans.

On the other hand, the width and the axial location of the shroud with respect to the impeller have
a significant influence on the air performance of a propeller fan (Guedel 1995a). A shroud that is
too wide (wider than 0.3 to 0.4 tip axial chord length), induces a deterioration of the fan curve.

The influence of the axial position of the impeller within the shroud is shown in Figure 25. When
the shroud is close to the blade trailing edge, the air flow at maximum free delivery is higher, but
the maximum fan pressure is lower than at the other shroud locations. An opposite trend is
observed when the shroud is near the leading edge. The influence of the shroud position on the
noise level is not very important. The shape of the shroud can be optimized to improve the
performance; for instance, by the use of CFD simulation (Guedel 1995a).

shroud Figure 27
impeller
Influence of the Axial Location of the
Shroud on the Fan Performance
Flow (Guedel 1995)
direction
motor 50

no. 1
40
Pressure (Pa)

x
+
30 + x

x
+
20
x
+
x +
no. 2 x
10
3 x 2 + 1

x
0 +

x
-10
0 500 1000 1500 2000 2500

no. 3 Flow rate (m3/h)

39
Fan Noise

1.3.2. Centrifugal fans

1.3.2.1 Introduction

There are three main types of centrifugal fans:

-backward-curved blades
-radial blades
-forward-curved blades

For these three fans, we shall present the main geometric and air performance characteristics, the
sound sources and their prediction, and some noise control methods.

1.3.2.2 Fan design and air performance

Backward-curved blade fans have blades that are curved or inclined in the direction opposite to
the rotation (Figure 28). Blades are of airfoil type or of constant thickness; airfoil blades being
slightly more efficient than the others. The number of blades ranges from 8 to 16. The aspect ratio
(inlet/outlet impeller diameter ratio) affects the fan pressure rise, while the blade width has a direct
influence on the volume flow.

Figure 28
Centrifugal Impeller with
Backward-Curved and
Backward-Inclined uniform thickness
Blades

aerofoil

A centrifugal impeller of small aspect ratio (less than 0.4) with narrow blades is able to provide high
pressure rise (10 kPa or more) at a rather small volume flow. Conversely, a wide blade impeller of
high aspect ratio (higher than 0.65) gives more flow, but its maximum pressure is smaller.

40
Parameters of Influence, Modeling, and Control of Fan Noise

Figure 29
Centrifugal Fan with a Scroll Housing ducts

The impeller is usually included in a volute of logarithmic shape with a discharge opening
perpendicular to the fan axis (Figure 29). Fan total efficiency may reach 80 to 90 % or more for this
configuration.

In some applications, the impeller is not installed in a volute, but in a concentric annular casing with
outlet guide vanes to convert the swirl velocity pressure into fan static pressure (Figure 30).
Performance of in-line centrifugal fans is slightly inferior to the classical centrifugal fans with volute.

Radial blade fans, with blades either completely radial or radial at the tip only (Figure 31), are used
in industrial applications where the handled air is corrosive or contains dust, sand, chips, etc. Fan
pressure may be as high as that of backward-curved fans, but the efficiency is usually lower than
the other two types of centrifugal fans, especially when the blades are fully radial (Daly 1992). The
number of blades ranges between 6 and 12, and the aspect ratio is similar to that of backward-
curved fans.

Figure 30
In-Line Centrifugal Fan

41
Fan Noise

Figure 31 radial tipped


Centrifugal Impeller with
Radial and Radial Blades
radial

Forward-curved blade fans (also called forward-curved fans with multi-vane impeller or squirrel
cage fans) are used in many applications: low pressure ventilation, air conditioning, heating,
household appliances, etc. The impeller has many blades (24 to 60, depending on the impeller
diameter) of small chord length, constant thickness, and curved in the direction of rotation (Figure
32). The maximum fan pressure is about 1000 Pa, and can be raised to 1500 Pa if the blades are
strengthened (Daly 1992). Fan efficiency does not exceed 60 to 70 %. We shall come back to the
flow pattern and performance of this type of fan in the next section.

Figure 32
Centrifugal Impeller with
Forward-Curved Blades

1.3.2.3 Description and prediction of sound generation mechanisms

The origins of noise for centrifugal fans are similar to those of axial fans. This section describes
the main parameters that affect the noise generation of backward-curved (BC) and radial
centrifugal fans; forward-curved (FC) fans are dealt with spearately since their noise sources are
different from the two other fans.

Tonal noise

Similar to axial fans, the two main causes of centrifugal fan tonal noise are the non-homogeneity
of the mean flowfield at the fan inlet and the interaction of the blade wakes with downstream
obstacles; i.e., scroll cutoff for fans with volute, and outlet guide vanes for in-line centrifugal fans.

An inhomogeneous mean flowfield at the fan inlet generates periodic loading on the blades which
radiate tones at harmonics of BPF. At the rotor outlet, the rotating wake of each blade is intercepted
by the volute cutoff. The cutoff plays the same role in noise generation as a single guide vane plays
at the discharge of an axial impeller: the scroll cutoff radiates noise at harmonics of BPF because
of fluctuating loads generated by the blade wakes on this part of the volute close to the impeller.

42
Parameters of Influence, Modeling, and Control of Fan Noise

Noise radiated by the cutoff usually prevails over the other tonal noise sources, especially when
the radial clearance between the blade tip and the cutoff is small (small clearance often being
necessary for aerodynamic reason). This can be explained by the velocity deficit in the blade wake,
which has more influence on noise close to the blade trailing edge than further away.

For this reason, the amplitude of the tones decreases significantly when the radial gap between
the impeller and the cutoff is increased, especially when the original gap is small.

Tonal noise is much more pronounced with radial fans, and, to a lesser extent, with backward-
curved fans than with forward-curved fans. This is easily explained by the large number of small
blades on FC fans, which induces less marked wakes than the two other impellers.

Equation 17, used for noise modeling of axial fans, also applies to the prediction of tonal noise of
centrifugal fans with a non-uniform mean velocity field at the fan entrance (for instance, this non-
uniformity may be due to the presence of an elbow at the fan inlet or to the wakes of inlet guide
vanes). Unlike axial fans, a centrifugal fan's axial component of the periodic blade load is
negligible, and the radial component of the load has to be taken into account in addition to the
tangential component to give a correct noise prediction (Raffaitin 1995).

An expression similar to Equation 18 can be adapted to predict the noise level radiated by the
volute cutoff. Experimental and theoretical works have been carried out for a better understanding
and prediction of the noise due to the blade wake-cutoff interaction (Thompson, Hourigan and
Stokes 1992, 197-204; Croba and Kueny 1992, 221-228; Leze et al. 1992, 213-220).

Broadband noise

Broadband noise mechanisms of centrifugal fans are similar to those of axial fans. The main noise
sources are:

-interaction of the inlet turbulent flow with the impeller


-interaction of the flow at the impeller exit with the volute cutoff, or, on in-line centrifugal fans, with
the outlet guide vanes
-blade self noise, which includes blade trailing-edge noise and vortex shedding noise.

The first two mechanisms may be described and predicted in the same way as what was done for
axial fans. This section focuses on blade self noise, even though this noise source has already
been discussed in the section relating to axial fans.

In a study of twelve centrifugal impellers, two conclusions were reached about blade design: it is
an important factor on the occurrence of flow separation on the blade, and the consequences it
has on broadband noise level (Bommes et al. 1995, 91-101). As a result of the study, a very simple
model to predict the relative increase of noise with flow separation was suggested.

Figure 33 shows three of the impellers tested. They have the same inner and outer diameters, the
same number of blades (B = 10), and the same leading edge angle, β1, but the blades are curved
in the backward, radial, and forward direction, respectively. The forward-curved impeller of Figure

43
Fan Noise

33 is not a multi-vane FC fan as drawn in Figure 32 and described in the next section.

In this study, the cutoff clearance is large (equal to 0.125 D) to minimize interaction noise and
emphasize blade self noise. Figure 34 shows the theoretical velocity vectors at the impeller exit for
the three blade curvatures. Relative flow velocity, w, which depends on blade angle, β, is known
along radius, r.

Figure 33
Blade Curvatures of Three 722 mm Centrifugal
Impellers (Bommes et. al 1995, 91-101)
β2 = 160°

R
39 93

β 2=
0.8
°

90°
2 11
1,

11 45
2 4

R R
β =

β 1=
β 1=

β 1=

132.5

31°
31°

31°

5 5 5
31 31 31
Φ Φ Φ
.3
83

2
2

72
72

R3

72

Φ
Φ

(a) (b) (c) (d) 181

The authors of the study then define convective acceleration, A, as:


∂w
A=w sin β
∂r

This term accounts for the flow deceleration across the impeller. The authors show a good
correlation between the fan broadband noise level and the convective acceleration at the blade
leading edge. When A reaches a given limit, flow separation occurs on the blades and broadband
noise level strongly increases. In the present study, this limit of A is reached on radial and FC fans,
but not on BC fans, which are therefore the quietest fans.

Burgain presents results of a detailed study on self-noise of in-line centrifugal fans; his conclusions
may be extended to other types of fans, i.e., axial fans (1998). The objective of this research,
essentially of experimental nature, was to correlate fan broadband noise to blade design and flow
characteristics of the blades. Systematic tests were made on fixed isolated profiles at different
angles of attack in an anechoic wind tunnel in order to examine the following mechanisms:

-noise due to the interaction of laminar boundary-layer instabilities with airfoil trailing edge
-turbulent trailing-edge noise
-vortex shedding noise.

44
Parameters of Influence, Modeling, and Control of Fan Noise

Figure 34
Velocity Vectors at the Impeller Exit for the Three
Blade Curvatures (Bommes et. al 1995, 91-101)

C2u C2u C2u

C C C2
C2m
2 2

C2m

C2m
W2 β2 W2 β2 W2
β2
U2 U2 U2

(a) (b) (c)

Noise due to laminar boundary layer instabilities (or Tolmien-Schlichting waves) occurs at low
angles of attack when the airfoil trailing edge is sharp and when the boundary layer on one or both
sides of the blade is laminar. The resulting radiated sound is a high frequency narrowband noise
centered on a frequency that varies as flow velocity to power 1.5. This noise disappears once the
boundary layer is turbulent.

Vortex shedding noise also happens at low angles of attack, but this mechanism occurs when the
airfoil trailing edge is thick compared to the boundary layer thickness. A narrowband noise results,
centered on frequency, f, such that f ≈ 0.2U/e, where U is the flow velocity and e is the airfoil
thickness at trailing edge.

Turbulent trailing-edge noise is the major self-noise source. It spreads in the medium to high
frequency range, with a shift towards the lower frequencies when flow separation occurs.

Characteristics specific to forward-curved blade fans

The flow pattern of a double-inlet, double-width, FC centrifugal fan is shown in Figure 35. Large
separation occurs at each inlet of the impeller, with a nearly axial flow field in the first quarter of
the wheel. This flow separation generates two contra-rotary vortices in the volute, which are clearly
seen at the fan discharge. There is only one vortex in a single-inlet configuration. The air flow
within the impeller is very unstable and disorganized in the cutoff sector, while it is well organized
and stable in the area diametrically opposed to the cutoff (Guedel 1995b).

A number of studies have been carried out to determine the influence of impeller and volute design
on air and sound performance of FC fans (Guedel 1995b; Morinushi 1987, 227-234; Konieczny

45
Fan Noise

and Bolton, 103-127). This research has allowed definition of design limits of the impeller and
volute that achieve good performance and low-noise level. Information on this issue is given in
Annex C.

With the constant improvement of the commercial CFD codes, it is now possible to predict the
performance curve of a FC centrifugal fan and the evolution of its performance with impeller and/or
volute geometry fairly accurately (Perrin 1997). Numerical simulations can progressively replace
expensive and time-consuming experimental studies carried out to assess the influence of
geometrical parameters on performance.

Figure 35
flow separation
Flow Pattern in a Double Inlet
Forward Curved Centrifugal Fan
(Guedel 1995b)

impeller axis

fan inlet

impeller

swirl

Figure 36 shows flow-pressure curves measured on six FC fans of diameters ranging from 118 to
140 mm. The curves are presented in non-dimensional coordinates in order to compare the
performance of these fans at the same speed and same diameter. Non-dimensional flow and
pressure coefficients are defined as:

qv
Flow coefficient: δ =
US

46
Parameters of Influence, Modeling, and Control of Fan Noise

Δps
Static pressure coefficient: μ =
pU 2

Where:

U = πDN
S = πDL (L is the blade width)

Discrepancies observed between these curves are due to differences in geometry between the six
fans. The shape and the number of blades, as well as the shape of the volute, play a significant
role on the performance (Guedel 1995b). Annex C provides information for the design of FC
centrifugal fans as a function of the operating point.

1.2
A
1
B
Pressure Coefficient, μ

C
0.8
D
E
0.6
F
0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6
Flow Coefficient, δ

Figure 36 - Performance Curves of Forward Curved Centrifugal


Fans Using Flow and Pressure Coefficients (Guedel 1995)

Figure 37 shows the variation of the specific noise level of nine FC centrifugal fans with respect to
the throttling coefficient Φ = δ / √2μ . The specific noise level is the overall A-weighted sound power
level radiated by both the inlet and outlet and converted to the following arbitrary reference
conditions: N = 1000 rpm, D = 100 mm, L = 100 mm, where D is the outer diameter and L is the
impeller width.

Acoustic similarity laws for FC centrifugal fans are different from those of axial-flow fans (see
Equation 8). For FC fans, the following conversion equations apply:

47
Fan Noise

N2 D L BW2 (f2 )
Lw2 (f2 ) = Lw1(f1 ) + 40 lg + 50 lg 2 + 10 lg 2 + 10 lg
N1 D1 L1 BW1(f1 )
Equation 19

N2
f2 = f1
N1

Where the subscript 1 refers to test data and the subscript 2 refers to scaled data. As in Equation
8, BW is the bandwidth of one-third band centered on frequency, f.

Notice in Figure 37 that the sound level continuously increases with Φ, i.e., when the point of
operation goes towards the higher volume flow on the fan curve. Unlike the other types of fans, the
noise level for this fan is not the minimum at the best efficiency point, for Φ = 0.25 to 0.3. Noise is
at a minimum at very low flow, however, it increases again as a blocked tight condition is
approached, i.e., in a range where the flow within the fan is much disturbed. Figure 37 also shows
that for a given value of Φ (i.e. for a given system characteristic on the non-dimensional
performance curve) the specific sound level does not vary significantly with the fan geometry,
provided that the blade-cutoff clearance is large enough.

60

55
LWA (dBA)

50

45
A B C D E F G H I
40

35
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Φ
Figure 37 - Specific Noise Levels of Forward Curved Centrifugal Fans.
Reference Conditions: N = 1000 RPM, D = 100 mm, L = 100 mm

The shape of both the impeller and the volute affects the air performance of a FC fan much more
than its sound level.

48
Parameters of Influence, Modeling, and Control of Fan Noise

1.3.2.4 Noise reduction methods Photo courtesy of Greenheck Fan Corporation

Tonal noise

As stated above, tonal noise of centrifugal fans is mainly due to the interaction of the blade wakes
with the volute cutoff. Different means have been successfully used to reduce this interaction
noise. These means, some of which are effective at controlling tonal as well as broadband noise,
are described in a review paper from Neise (1982, 151-161).

• Increase of the cutoff clearance: the amplitude of the BPF tone continuously decreases when
the cutoff clearance increases. An increase of the clearance from 0.7 to 12% of the impeller
diameter induces a 20 dB reduction of the BPF level of a BC centrifugal fan. This method also
works on FC fans, and it results in reduction of both the tonal and the broadband noise level
(Konieczny and Bolton 1995, 103-127). The minimum clearance recommended for fan noise
control is 10 to 12% of the impeller diameter. Such a clearance does not reduce fan performance.

• Increased radius of curvature of the cutoff: this parameter has been studied on a BC fan. An
increase of the cutoff radius of curvature from 0.5 to 10% of the impeller diameter results in a 6 dB
noise reduction without deterioration of the fan performance.

• Angle between the cutoff edge and the trailing edge of the impeller blades: on a
conventional centrifugal fan, the cutoff is parallel to the blade trailing edge. The sources of
interaction noise distributed on the cutoff radiate in-phase, and the radiated sound level is
therefore at a maximum. A cutoff inclination angle of one blade spacing, which introduces a phase
difference between the sources, reduces the amplitude of the tones, especially when the cutoff
clearance is small.

• Circumferential shift between the wheels of double-inlet, double-width fans: if the blades of
the first wheel lie between the blades of the second wheel, the BPF noise can be reduced by 8 to
10 dB.

49
Fan Noise

• Uneven blade spacing: this has already been mentioned for axial fans and the same conclusion
can be drawn: this method is valid to reduce amplitude of the BPF harmonics, but the initial
acoustic energy is spread over other harmonics of the shaft frequency, and the overall sound level
remains unchanged.

• Acoustic resonator in the cutoff: the use of a quarter wavelength resonator with its opening
flush-mounted on the cutoff wall can considerably reduce the BPF tone level. This reduction is
achieved by adapting the resonator length to the frequency that has to be eliminated. To reduce
both the fundamental and the first harmonic of BPF, a device with two cavities has been suggested
and validated (Tournoy and Bruyere 1992, 279-285).

• Active noise control (ANC): ANC at the source, with an active source (e.g., loudspeaker) in the
cutoff, is a very efficient means of BPF noise control, even if the shaft frequency changes with time
(Neise and Koopmann 1988, 209-200; Besombes 1993). This ANC system is different from the one
classically used in a ventilation duct. The latter is efficient for controlling tonal and broadband noise
at low frequency, such as in the plane wave propagation range, while the former works for reducing
tonal noise only, but in a wider frequency range below and above the first duct cutoff frequency (for
a cylindrical duct of diameter, D (m), the first cutoff frequency is f ≈ 200 / D).

Photo courtesy of Howden Buffalo, Inc.

50
Parameters of Influence, Modeling, and Control of Fan Noise

Broadband noise

In addition to the methods mentioned previously, other noise control means are specific to
broadband noise, although they may be effective, to some extent, for tonal noise reduction (Neise
1982, 151-161; Riera-Ubiergo 1988).

• Choice of the fan operating point: on BC centrifugal fans, the noise level is minimized at its
best efficiency point (BEP). It is therefore recommended to operate the fan at BEP.

Conversely, as seen above on FC fans, the noise level is not minimized at BEP but at a volume
flow closer to shut off condition. If a low sound level is required, it is recommended that the fan be
operated at a throttling coefficient, Φ = 0.05 to 0.1 (see Figure 37). Coefficient Φ can be rewritten
as:

⎛ 1⎞ qv
Φ =⎜ ⎟
⎝S ⎠ Δp
2
ρ

where: S = πDL (D is the outer diameter and L is the width of the impeller).

This means that to reach a low value of Φ at given qv and Δp, the impeller area, S, must be as
large as possible. This is an additional argument in favor of a large impeller instead of a small one
for noise reduction.

• Radial clearance between the inlet cone and the impeller: on BC fans, the inlet cone reaches
into the impeller mouth with an axial overlap of a few millimeters. The radial gap between the cone
and the impeller in this overlap plays a significant role in the aerodynamic and acoustic
performance. This gap has to be as small as possible.

On FC fans, the inlet nozzle diameter is nearly equal to the wheel inner diameter and does not
extend into the wheel. The axial clearance between the nozzle and the impeller is of the order of
3 to 5 mm, and a reduction of this clearance does not improve fan performance, nor reduce noise.

• Optimization of the volute shape: the shape of the volute has a significant effect on the airflow
performance; more so than on the noise level, if cutoff clearance is accepted. It is, therefore,
important to optimize the volute shape to achieve the required operating point at the lowest
possible speed. This optimization may be done by parametric tests of various volute models, or,
as are becoming more and more common, by parametric CFD calculations.

On FC fans, a tightened volute is well adapted to low flow rate, while a flared volute is preferable
at higher flow rate (Guedel 1995b; Konieczny and Bolton 1995, 103-127). See also Annex C.

• Acoustic lining in the volute: an acoustic lining on the inner wall of the volute reduces tonal and
broadband noise at the fan discharge. However, the noise radiated by the inlet is not attenuated
(Neise 1982, 151-161; Szentmartony, Kurutz, Koscso, 271-278).

51
Fan Noise

• Reduction of blade self noise: methods mentioned in Section 1.3.1 to reduce blade self noise
of axial-flow fans also apply to centrifugal fans. For example, it is advisable to avoid flow separation
on the blade by running at the optimum operating point (except on FC fans where blade self noise
is not a predominant mechanism), and to eliminate potential vortex shedding noise by beveling the
blade trailing edge.

Burgain reports the appearance of laminar boundary layer instabilities on the blade pressure side
of an in-line BC centrifugal fan with the emergence of a strong narrowband noise centered on 3
kHz (1988). The author succeeds in eliminating this hump by mounting a mesh screen in the fan
inlet to make the blade boundary layers turbulent. This does not degrade fan performance nor
increase the low frequency interaction noise due to the augmented inflow turbulence.

Another example is reported where the inner and outer circumference of a FC impeller is covered
by mesh screens (Figure 38) to generate a more stabilized turbulent boundary layer on the impeller
blades and to create a smoother velocity field at the blade trailing edge (Neise 1982, 151-161).
Broadband noise level is strongly reduced, but it is to the detriment of fan performance in this case.

meshes
Figure 38
Mesh Screens on the Inlet and Outlet of a
Foward Curved Centrifugal Impeller
(Neise 1982, 151-161)

In one of his prior studies, the author of this book observed a very strong whistle at 9 kHz on a FC
centrifugal fan mock-up. This noise was likely due to laminar boundary layer instabilities, and it
disappeared when a single strut was put across the fan inlet, acting as a turbulence generator.

Solutions already mentioned for the noise control of axial flow fans, such as saw-tooth or beveled
blade trailing edge (see Figures 25 and 26), may also be adopted to reduce blade self noise of
centrifugal fans.

1.3.3. Mixed-flow fans

Mixed-flow fans are used much less commonly than axial and centrifugal fans, therefore, acoustic
studies on these fans are less numerous than on the two other types of fans. Sound generation
mechanisms of mixed-flow fans are very similar to those of axial-flow fans. Trends of noise levels

52
Parameters of Influence, Modeling, and Control of Fan Noise

observed on vaneaxial fans, such as the effect of rotor-stator axial clearance, number of blades
and vanes, and blade tip clearance are also found on mixed-flow fans (Ameziane, Pauzin, and
Guilhot 1992, 179-186; Ameziane 1992; Fukano and Kodama 1992, 105-112).

Prediction of tonal noise is improved when the radial component of the unsteady blade load is
taken into account. Comparisons of predicted and measured tonal noise levels generated by the
outlet guide vanes have been studied (Ameziane, Pauzin, and Guilhot 1992, 179-186; Ameziane
1992).

The specific noise levels (noise levels at the same fan performance) of mixed-flow fans are slightly
lower than those of axial fans (Ameziane 1992; Neise 1992, 45-56).

1.3.4. Cross-flow fans

1.3.4.1 Introduction

The principle of the cross-flow fan was discovered more than one century ago, and its original
application, like most types of fans, was mine ventilation. Despite its old origin, this fan was not
studied or improved much until recently. A renewed interest in this fan appeared with the expanded
use of small air-conditioning units and household appliances, such as domestic ovens. Its easy
integration into the appliance provides a size advantage in comparison to other fans with the same
performance.

Its working principle, the geometrical parameters that influence air and sound performance, and a
few means of noise control, are presented in the following sections.

1.3.4.2 Flow pattern in a cross-flow fan

Studies carried out to gain a better understanding of the flow pattern in this fan, and for defining
the parameters that influence the performance have been done by different researchers, including
Eck (1973). Until recently, these studies were essentially based on experiment, such as flow
visualization and velocity measurements inside and outside of the impeller. Currently, CFD
calculations more frequently replace experiment when predicting mean flow pattern and fan
performance. Also, there is a noticable time reduction when a parametric study is carried out with
the aim of optimizing the fan geometry.

A cross-flow impeller has many forward-curved blades of small chord. The geometry of the impeller
is very similar to that of a FC centrifugal fan, except it is much wider. The fixed part of the fan (or
stator) is made up of a tongue and a volute (Figure 39) that strongly influence the flow pattern and
the performance.

Unlike a centrifugal fan, the flow pattern for a cross-flow fan is two-dimensional, with flow lines
remaining in a section perpendicular to the impeller axis. Fan function is controlled by a vortex
whose position in the impeller depends on the operating point: the vortex extends across the
blades and is located close to the tongue at large volume flow (Figure 39), while it moves towards
the impeller axis at low volume flow. The tongue mainly ensures the stability of the vortex. A small

53
Fan Noise

part of the flow goes through the gap between the impeller and the tongue to feed the vortex, as
shown in Figure 40, where a result of a CFD flow simulation in a cross-flow fan is presented
(Gautier et. al 1998). This type of fan is very unstable at low flow rate because of the unstable
location of the vortex within the impeller in this operating range.

Figure 39
Flow Pattern in a Cross-Flow Fan

impeller

tongue

volute

Figure 40
Flow Simulation in a Cross-Flow Fan

1.3.4.3 Influence of fan geometry on aerodynamic and acoustic performance

The aerodynamic and acoustic performance of a cross-flow fan depends not only on the geometry
of the fan (impeller and stator), but also on the surroundings of the inlet and outlet. This last point
concerns fan installation effect, which is dealt with in Section 2. Unlike the other fans, we consider
system effect of cross-flow fans separately in this section.

Studies dealing with the optimization of cross-flow impellers are not numerous, but useful trends
can be learned from the results of this work (Gautier et al. 1998). As dicussed previously, the
impeller is very similar to that of a FC centrifugal fan, and the specifications made in Section 1.3.2
for FC fans also apply to cross-flow fans, with the exception that the impeller width may be very

54
Parameters of Influence, Modeling, and Control of Fan Noise

large (8 to 10 wheel diameters). This width is only limited by mechanical constraint.

With the same symbols as those used for FC fans (see Annex C), the design of a low-noise, cross-
flow impeller should adhere the following specifications:

-inner/outer diameter, D1/D2 ≈ 0.8


-solidity, C/t ≈ 1
-chord length/outer diameter, C/D2 ≈ 0.1
-blade pitch angle, γ = 35° to 40°
-blade camber angle, δ = 70° to 75°.

The stator also has a significant influence on the performance:

-As mentioned above, the cut-off has a major role in stabilizing the vortex, and therefore, in the fan
operation as well. As the cut-off clearance is reduced, fan performance improves, but fan noise
increases. A reduction of the clearance induces an increase of the BPF peak, however, the
broadband noise level is not much affected by this variation in gap (Hofe and Thien 1992, 237-244;
Guedel and Freynet 1996). An optimum clearance may be found, which leads to a minimum
specific sound level (see below).

-The cut-off clearance must not be larger than the scroll clearance, which is the smallest gap
between the impeller and the scroll. If this condition is not fulfilled, the fan curve could be
significantly degraded. The cut-off shape also has an influence on performance (see below).

-The shape of the scroll, as well as the angular positions of the cut-off and scroll cut-off, also have
a great influence on performance (see below).

The variation in sound level of the cross-flow fan with operating points is similar to what is observed
on FC centrifugal fans (see Section 1.3.2): the noise level continuously increases with air flow. The
control of rotational noise is more efficient at high flow because the tonal noise level usually
increases more quickly with the flow than the broadband noise.

Similar to other types of fans, the impeller diameter has noticeable influence on the performance.
Due to Reynolds effects, fan laws do not strictly apply, and the impellers with large diameter are
more efficient. Experimental results on two geometrically similar fans of 60 and 100 mm impeller
diameters show that for the same point of operation, the small fan is significantly noisier than the
large one, mainly due to poor aerodynamic performance (Guedel, Freynet, and Boiteux 1996).

Fan inlet and outlet surroundings have significant influence on the flow pattern in the fan, and
therefore, on the fan performance. Figure 41, taken from a Japanese fan manufacturer catalog,
shows good and bad examples of cross-flow fan installation (crosses = bad installations, rounds =
good installations). The hatched area on the bottom drawing should be free of walls or obstacles.

It is advisable to fit the fan outlet with a diffuser to convert a part of the velocity pressure into static
pressure. In some applications, the fan is coupled with a heat exchanger, which plays the same
role as a diffuser when it is placed at the fan discharge (Hofe and Thien 1992, 237-244).

55
Fan Noise

The optimization of the fan surroundings can be made either by test and flow visualization or,
recently, by CFD calculation.

Aerodynamic and acoustic fan laws of cross-flow fans are the same as those of FC centrifugal
fans. The definition of non-dimensional flow and pressure coefficients δ, μ, as well as conversion
equation of sound levels, are therefore identical to those used for FC fans in Section 1.3.2.

X O D O D

X O D
O
D

X O O

X O X O

X O 1.5D

1.5D

Figure 41 - Guideline for Cross-Flow Fan Installation


(from Royal Electric Fan Manufacturer Catalog)

1.3.4.4 Noise reduction methods

The main source of tonal and broadband noise of cross-flow fans is the interaction of the blade
wakes with the cut-off and scroll entry cut-off. The following measures can minimize this interaction
noise:

• increase the cut-off clearance: the radial clearance between the impeller and the cut-off must
be large enough to have a low noise level. Unlike FC fans, cutoff clearance on a cross-flow fan has
a strong effect on fan performance. If the clearance is too large, the fan curve degrades and fan
performance is significantly reduced. Therefore, a compromise has to be found to minimize
specific sound level at a given operating point. For instance, in a parametric experimental study by

56
Parameters of Influence, Modeling, and Control of Fan Noise

Guedel and Freynet, it is shown that the best compromise for the cutoff clearance is 6% of the
impeller diameter at low flow rate in a range where the BPF spike does not protrude very much
above the broadband noise, and 10% of the diameter in the high flow region (1996).

• spanwise twist of the impeller: twisting the impeller can control noise significantly, especially
when the cut-off clearance is small, in order to keep good fan performance. The reason is to allow
the sources distributed along the cutoff to radiate the sound with a phase lag. A similar method is
to build an impeller with several sections; each section being angularly shifted from the adjacent
ones.

• uneven blade spacing: the same as noise control of axial and centrifugal fans. The BPF tones
are reduced, but peaks at multiples of the shaft speed emerge.

• optimization of the cut-off geometry: results of a parametric study on an impeller of 100 mm


in diameter show that a fan with a cut-off made from a sheet of 1.5 mm thickness (Figure 42) has
high performance, especially if its cutoff clearance is small (6% of the impeller diameter or less)
(Guedel and Freynet 1996). The downside is that the sound level is very high, mainly because of
a very strong peak at BPF. The fan specific sound level is still quite high in comparison with other
cut-off shapes because its high aerodynamic performance and the reduction of speed to achieve
a specified operating point cannot compensate for the loud radiated noise.

A more interesting cut-off geometry, as far as noise level reduction is concerned, is shown in Figure
43. The cut-off shape is slightly less aerodynamically efficient than the cut-off in Figure 42, but it is
considerably less noisy. The porosity of the perforated sheet used was varied from 23% to 50%
and did not have much influence on the aerodynamic and acoustic performance. A perforated
sheet is slightly better than a solid sheet of the same shape for unknown reasons. It is possible
that a porous cutoff, in which a small amount of air flow passes, has a positive influence on the
vortex feeding and stabilization.

Figure 42
Tongue Made Up of a Blade of 1.5
mm Thickness
(Guedel and Freynet 1996) ~135°

blade of 1.5 mm thickness

57
Fan Noise

Figure 43
Tongue Made Up of a Curved
Perforated Sheet solid sheet
(Guedel and Freynet 1996)

perforated sheet

• optimization of the scroll shape: the geometry of the scroll has an influence as shown in
(Guedel and Freynet 1996). Out of the three scrolls that have been tested, the one that yields the
lowest specific noise level has the following geometry:

-scroll expansion: a = 0.006, where a is defined by: r(θ) = r0 (1 + aθ), and where r is the
radius of the scroll at angle, θ, and, r0, the radius at scroll cutoff. θ is in degrees, and θ = 0° at
the cutoff.

-scroll clearance: 10% of the impeller diameter.

-cut-off clearance: 6 to 10% of the impeller diameter, depending on the operating point.

-angular positions of the scroll and cutoff: the two cutoffs are diametrically opposed.

• guide vane in the impeller: a fixed guide vane was successfully applied within the impeller of a
cross-flow fan for automotive engine cooling (Figure 44) (Hofe and Thien 1987, 3.80-3.89). The
guide vane, which spreads over the whole impeller width, is to optimize the shape and position of
the vortex in the impeller. According to the authors, a significant improvement of the fan
performance occurs, without a noise increase, due to a reduction of the aerodynamic losses. The
specific noise level of the configuration with the guide vane was therefore significantly lower than
the initial configuration without the guide vane.

Figure 44
Flow Patter in a Cross-Flow Fan with
a Guide Vane in the Impeller guide vane
(Hofe and Thien 1987, 3.80-3.89) vortex
tongue

diffuser

58
Parameters of Influence, Modeling, and Control of Fan Noise

1.4. Concluding Remarks

The mechanisms of aerodynamically generated noise are reasonably well understood.


Aerodynamic noise sources of dipole type are created by the periodic and random fluctuating loads
on rotating blades and fixed obstacles such as struts, guide vanes, or scroll cutoff. Periodic loads
generate tones, while random loads generate broadband noise.

Broadband noise can be split up into self-noise, which results from the passage of the flow on the
blades without any external disturbances, or interaction noise, which depends on inflow turbulence
and fixed obstacles upstream and downstream of the impeller.

Acoustic fan laws, which permit conversion of measured sound levels to other fan sizes and
speeds, are validated. Reliable fan noise testing methods now exist since the publication of the
international standard ISO 13347.

Shortages exist in fan noise analysis, prediction and control methods, especially for broadband
noise, which is often the major contribution to the overall A-weighted sound level. To achieve
significant fan noise reduction requires finding the actual sound generation mechanisms and
sound radiation. Based on that knowledge, sound models can be used to relate the geometrical
fan parameters and the flow patterns in the fan point of view of emitted sound field.

A universal theoretical method that could be implemented and would allow the prediction of noise
characteristics of any fan geometry for any flow condition does not exist. Computational Fluid
Dynamics (CFD) now allows researchers to predict fan curves with reasonable accuracy. For noise
prediction, it is more relevant to implement and to validate models adapted to each identified sound
mechanism. Some of them have been described in this chapter. It is not expected that these
models will predict fan noise levels very accurately, but if they are intended to be used for low-noise
fan design, they must provide the correct trends when the geometric and flow parameters are
varied.

One of the difficulties in usage of these prediction methods is to get suitable input data for the
models. For instance, the input data of the trailing-edge noise prediction model are the spectrum
and spanwise correlation length of the wall-pressure fluctuations at the blade trailing edge. These
data cannot yet be obtained from classical CFD calculations using Reynolds Average Navier-
Stokes (RANS) model. More advanced numerical techniques like Large Eddy Simulation (LES) or
Direct Numerical Simulation (DNS) could provide these data in the future.

Experiments with flush-mounted fluctuating pressure probes on rotating blades are possible (some
authors carried out measurements with a few rotating probes), but they are still expensive and not
reliable for at least two reasons: signal transmission and probe miniaturization. Furthermore,
several probes distributed spanwise on the trailing edge are needed to get the wall-pressure
coherence scales, which is expensive and sometimes impossible to do on a rotating blade.
Conversely, unsteady pressure measurements on fixed airfoil in a wind tunnel are much easier to
do and may give information that can be extrapolated to rotating blades. Flow velocity
measurement between rotating blades by Laser Doppler Velocimeter (LDV) or Particle Image
Velocimeter (PIV) may also provide useful data. Some measurements with hot wires rotating on

59
Fan Noise

axial impellers were also made. An important problem when using this technique is signal
transmission, which is done either by remote transmission or by the use of a slip ring.

All of the above methods, either theoretical or experimental, are for the most part difficult to apply
to industrial fans. Therefore, it is worthwhile to develop simplified models in order to draw useful
trends for the understanding, prediction and control of fan noise mechanisms.

60
2 Fan Noise Installation Effects

2.1 Definitions and examples

2.1.1 Introduction

Fan installation effects, or system effects, are a significant reason why studies are carried out on
fans because of their significance in the practical application of fans. When a fan is installed in a
system, its aerodynamic and/or acoustic performance may be quite different from what is expected
when compared to the performance of the fan alone. Application of the fan in a system often leads
to deterioration of fan performance. It is important to understand, predict and possibly reduce
system effects in order to choose the correct fan for the system at the system design stage.

If system effects are ignored, fan specifications (air flow, efficiency, noise level) are not likely to be
achieved. This section targets the acoustic installation effects, but a few words need to be said
about aerodynamic system effects first.

2.1.2 Aerodynamic installation effect

Aerodynamic fan system effects result in a deterioration of fan performance compared to a


standardized test installation, as shown in Figure 45. Without system effects, the airflow volume of
the fan in the system is Qv0 at the intersection of the fan and system curves. When an installation
effect occurs, the fan curve deteriorates and the corresponding volume flow is therefore Qv1, such
that Qv1 < Qv0.

AMCA Publication 201 defines the three main causes of system effects on fan performance (1990).
They are:

-non-uniform flow at the fan inlet


-swirl at the fan inlet
-flow distortion at the outlet

Besides degradation of fan performance due to installation effects, the decline may also be due to
the difference between the estimated and actual system curves. This may occur if the air velocity
is non-uniform at the fan discharge, resulting in an underestimation of the system pressure losses.

61
Fan Noise Installation Effects

Figure 45 Fan curves


Aerodynamic Fan Installation Effect Ps
Without system
effect

With system Duct system


effect curve

Qv1 Qv0 Volume flow

For different types of fans and duct configurations, AMCA Publication 201 gives a quantitative
estimation of the system effect factor (SEF). For example, the SEF due to an elbow at the fan inlet
or outlet is quantified according to the type of fan (tubeaxial, vaneaxial, centrifugal), shape and
orientation of the elbow, axial distance between the elbow and the fan, and flow velocity.
Recommendations are given in order to minimize this effect.

2.1.3 Acoustic installation effect

2.1.3.1 Definition and origins

Acoustic installation effects usually result in an increase of the sound power level of the fan
installed in the system when compared to the level of the fan tested alone at the same operating
point. This effect has two main causes:

-Deterioration of the inflow conditions due to non-uniformity of the air velocity field and/or an
increase in turbulence of airflow entering the impeller. For example, an obstacle or an elbow at the
fan inlet may cause this disturbance. A duct singularity at the fan discharge may also have some
effect on noise if it is very close to the impeller. See Section 2.2.

-Acoustic loading effect associated with the reflection of the sound waves radiated by the fan
into the duct system. This effect modifies the sound power level as determined by the fan alone in
a standardized test rig where the sound reflections are strongly attenuated by anechoic
terminations at the duct ends. See Section 2.3.

Other causes may explain differences in the sound power levels measured on-site and in
standardized installations, but they are not associated with system effects. Those causes are:

-sound attenuation in the system


-flow generated noise
-fan installation conditions

62
Definitions and Examples

The attenuation of the fan sound power in a system may be due to the absorption and reflection
of sound waves in the duct, as well as duct breakout noise. Sound attenuation in systems,
including lined or unlined straight ducts, elbows, abrupt duct transitions, branches, etc., has been
assessed experimentally by a number researchers (AICVF 1997). Acoustic system effects are not
involved in this case because the fan may be replaced by any other sound source (i.e. a
loudspeaker) without significant change in the noise attenuation.

Flow generated noise by straight ducts and duct singularities has been extensively studied
(AICVF 1997). Once again, this mechanism has nothing to do with fan system effects since it
results from an interaction of the air flow with duct components and not with the fan itself.

Fan installation conditions have been discussed in Section 1.2.3. Four standardized test
installation categories are defined in Figure 10. Fan sound power levels may differ considerably
from one installation category to another (Neise 1992, 45-56), and it is recommended that the
sound levels be determined according to the category that is the most similar to the on-site
installation. A difference measured in sound levels on a given fan mounted in a standardized test
rig and then mounted in an actual system cannot be viewed as a system effect if the fan
installations are different. For instance, a propeller fan mounted in an air-conditioning unit has to
be tested according to a Type A installation category. If the standardized test is not made in a Type
A installation, the potential difference between on-site and standardized results cannot be
attributed to a system effect.

2.1.3.2 Examples of acoustic system effects

A situation that often occurs in practice is the presence of a duct elbow at the fan inlet. Neise
reports a 14 dB increase in the low-frequency noise of an axial flow fan due to a 90° elbow near
the fan inlet, and a 4.5 dB(A) increase in the overall sound power level of a centrifugal fan with an
inlet box when compared to the level with an axisymmetric opening (Neise 1992, 45-56). Figure 46
shows a comparison of the sound power spectra measured at the inlet of a tubeaxial fan of
diameter D = 350 mm that is fitted with different inlet duct configurations (Guedel, Andre, and
Freynet 1998). The fan is connected either to a straight duct of 4D length (reference configuration),
or to a duct with a 90° elbow at various distances between (0 and 3D) to the fan inlet. Very
significant discrepancies are observed between these spectra: at low frequency, the reference
configuration is the quietest, while at 315 Hz and above, the duct with the bend at 3D is the
quietest.

For air-cooling and air-conditioning units, the fan, which is usually of axial-flow type, is placed in
front of a heat exchanger in a plenum chamber. The heat exchanger is usually placed at the fan
inlet, and seldom at the fan discharge. Acoustic system effects may be observed due to the flow
disturbance induced by the exchanger when it is at the inlet. Figure 47 shows an example of
system effects associated with the presence of a heat exchanger at the fan inlet of a small air
condenser. Three propeller fans with diameters of 360-mm and different geometry have been
tested in the same unit. For two of these fans, at approximately 400 Hz, the level is 7 dB higher
with the exchanger than without. For the third fan, the difference is much lower.

63
Fan Noise Installation Effects

90 BPF
Straight duct
85 Elbow at 0D
Elbow at 0.5D
80 Elbow at 3D
L W (dB) 75

70

65

60
50

100

200

400

800

1600

3150

6300
f(Hz)
Figure 46 - Influence of a 90º Elbow at the Inlet of a Tubeaxial Fan
On Its Inlet Sound Power Spectrum (D = 350 mm)

10
2 BPF
BPF
5

∆LW(dB) 0
BPF
2 BPF
-5
4 BPF ∆LW = LW fan + exchanger -LW fan alone
2 BPF
-10 100 Hz motor
f
63
80
100
125
160
200
250
315
400
500
630
800
1000
1250
1600
2000
2500
3150
4000
5000
6300
8000

(Hz)

Figure 47 - Acoustic System Effect Due to a Heat Exchanger Upstream of Three


Propeller Fans of 360-mm Diameter and Different Geometry (Guedel 1995)

64
Effect of Inflow Conditions

2.2 Effect of Inflow Conditions

2.2.1 Origins and main characteristics

Acoustic system effects associated with inflow conditions have already been mentioned in Section
2.1. The non-uniformity of the velocity field at the impeller inlet may increase the amplitude of the
periodic load on the blades, and therefore, the amplitude of the BPF discrete tones. The
broadband noise level may increase with an inflow turbulence rise.

Experience shows that axial fans are usually more subject to inflow condition effects than
centrifugal fans. Figures 48 and 49 show the influence of a 90° duct elbow on a tubeaxial fan and
a BC centrifugal fan installed at different distances from the inlet (Guedel, Andre, and Freynet
1998). Tests on the two fans are made for the same duct system. Figure 48 is in fact based on data
from Figure 46, where the difference in levels between the bent duct and the reference straight
duct was plotted against frequency for the various elbow-fan distances. A comparison of Figures
48 and 49 shows that the system effect, accounted for by ΔLW, is significantly higher for the axial
fan than for the centrifugal fan, where the fan values of ΔLW are still important at some frequencies
because of the acoustic impedance effect (see Section 2.3).

Another lesson that can be learned from Figure 48 is that the amplitude of the system effect
decreases when the inlet obstacle is moved away from the fan. This result is logical because the
flow at the impeller inlet is more uniform and less turbulent when the obstacle is far away.

20

15 elbow at 0D
elbow at 0.5D
10
elbow at 3D
5
∆L W (dB) 0
100

200

400

100

1600

3150

6300
50

-5
-10

-15
-20 f (Hz)

Figure 48 - Acoustic System Effect Due to a 90º Elbow at Various Distances From the
Inlet of a Tubeaxial Fan (Duct Diameter D = 350mm) ΔLW = LW with elbow - LW without elbow

65
Fan Noise Installation Effects

20

15 elbow at 0 D
elbow at 0.5 D
10
elbow at 3 D
5
∆L W (dB) 0
100

200

400

100

1600

3150

6300
50

-5
-10

-15
-20 f (Hz)

Figure 49 - Acoustic System Effect Due to a 90º Elbow at Various Distances


From the Inlet of a Backward-Curved Centrifugal Fan (D = 350 mm)

Noise system effects are also reduced when the obstacle is at the fan discharge, as shown in
Figure 50, which uses the same axial fan and the same duct elbow as in Figure 48. When the
elbow is very close to the fan outlet (0D or 0.5D distance), an effect due to the disturbance of the
flow in the impeller is noticeable, which induces a slight increase in the low-frequency noise
(Guedel, Andre, and Freynet 1998). A flow disturbance has a difficult time travelling upstream;
however, it is easily propagated downstream. This explains the influence of distance and
singularity location, with respect to the fan, on the amplitude of the system effect.

Guedel confirms this last remark with the results shown in Figures 51 and 52 (1995). These figures
show, respectively, values of system effects due to a heat exchanger located at the inlet and the
discharge of an axial-flow fan at a distance of about 0.5D. Four fans of diameters 710 to 760 mm
have been tested with and without the same heat exchanger with dimensions of 1190 x 1160 mm
in Figure 51. In both figures, the average system effect curves are also plotted as a bold
continuous line. The system effect is negligible, or there is even a "positive system effect", i.e., a
sound reduction in the 63 Hz octave band, when the coil is on the fan outlet side. Conversely, a 5
dB increase of the sound power level is observed at 125 Hz when the exchanger is upstream.

Experience also shows that some axial fans are more subject to inflow condition effects than
others. The amplitude of this effect depends on the impeller geometry in a way not yet fully
understood. In the examples of Figure 47, the fan that has the lowest system effect has four long
chord blades, while the two other fans have seven narrow short chord length blades.

66
Effect of Inflow Conditions

20

15 elbow at 0 D
elbow at 0.5 D
10
elbow at 3 D
5
∆L W (dB) 0 100

200

400

800

1600

3150

6300
50

-5
-10

-15
-20 f (Hz)

Figure 50 - Acoustic System Effect Due to a 90° Elbow at Various


Distances From the Outlet of a Tubeaxial Fan (D = 350 mm)

10
fan A
dLW = LW(with coil) - LW(without coil)
fan B
5 fan C
fan D
dLW (dB)

Average
0

-5

-10
63

125

250

500

1000

2000

4000

8000

f (Hz)
Figure 51 - Acoustic System Effect Due to a Heat Exchanger at the Inlet of Propeller
Fans of Different Geometry and Diameters Between 710 and 760 mm

67
Fan Noise Installation Effects

10
fan A
fan B
5 fan C
fan D
dLW (dB)

Average
0

-5

-10
63

125

250

500

1000

2000

4000

8000
f (Hz)
Figure 52 - Acoustic System Effect Due to a Heat Exchanger at the
Discharge of Propeller Fans (Same as Figure 51)

2.2.2 Description and prediction of the mechanisms

This section returns to some results and comments of Section 1.3.1 in order to continue the
analysis on the effect of inflow conditions.

2.2.2.1 Tonal noise

We have seen in Section 1.3.1 that the amplitude of the BPF tones could be predicted from
Equation 17. When the air velocity at the impeller inlet is non-uniform because of upstream
obstacle wakes, uneven flow field induced by an elbow, or anything else in the system, periodic
forces on the blades increase, which results in an augmentation of the tonal noise level.

The prediction of the inflow effect consists of the assessment of the increase in the blade loading.
As seen previously, this estimation may be made from the knowledge of the air velocity pattern in
a section close to the blade leading edge, and from an aerodynamic transfer function between the
inlet flow field and the blade forces. This approach may look simple, but it is still difficult to
implement if a higher accuracy is desired. One of the difficulties is obtaining the correct velocity
pattern in the inlet section of the rotating impeller. Experimental means such as Laser Doppler
Velocimeter (LDV) or CFD simulation may be useful to achieve these input data. The aerodynamic
transfer function is mentioned in Section 2.2.3.

2.2.2.2 Broadband noise

Broadband noise mechanisms related to inflow effect have already been mentioned in Section
1.3.1. This section details the influence different parameters have on both the interaction of the
turbulent inflow with blades and their role in fan noise increase. Blake provides useful information
to this point (1986).

68
Effect of Inflow Conditions

The sound power spectrum has a direct link to the blade lift’s random fluctuations, which depend
on the statistical characteristic of the turbulent inflow. The shape of the noise spectrum varies
according to the correlation length of the fluctuating velocities at the impeller entrance. If the scale
of the flow turbulence in the axial direction is smaller than the distance between two adjacent
blades, then the noise spectrum is of a broadband type, since the noise radiated by each blade is
uncorrelated from one blade to the other. On the other hand, if the scale of the flow turbulence is
much larger than the blade spacing, then the sound sources on the blades are more or less
correlated, and the fan noise spectrum shows spikes around harmonics of BPF.

If the structures in the flow are very elongated in the axial direction because of the suction effect
of the rotating machine, they are chopped by the blades while they reach the impeller. This
mechanism may strongly enhance the amplitude of the BPF tones, but it is mainly encountered on
high-speed fans in aeronautical applications.

The level and shape of the fan broadband spectrum depend of the following parameters:

-Turbulence intensity of the axial inlet velocity. The spectrum level increases as 20 log α,
where α is the ratio of the turbulence rates with and without inflow disturbance. If the turbulence
rate is doubled, the noise spectrum increases to 6 dB in the frequency range where this
mechanism occurs.

-Spanwise correlation length of the axial fluctuating velocity. An increase in the turbulent
scales leads to an acoustic energy transfer from high to low frequencies (Fournier 1988).

The mechanism of turbulence ingestion mainly affects the low and medium frequency part of the
noise spectrum (typically below about 1 kHz) while blade self-noise tends to prevail in the high
frequency range (see Section 1).

Analytical models have been implemented to predict the effect of inflow conditions on fan noise,
but comparisons of predicted and measured results are rather scarce (Blake 1986; Fournier 1988).
The input data to the models, namely the spectrum and spanwise coherence length of the
fluctuating velocity field at the impeller inlet, are difficult to obtain accurately by experiment since
the velocity measurements have to be made in a section close to the blade leading edge. If hot
wires are used, they cannot be placed close to the rotating impeller. LDV does not have the same
drawback, but at least two channels are required to get the spanwise length scales of the turbulent
velocity. CFD cannot yet provide these input data with enough accuracy.

Conversely, a number of studies have been conducted on the broadband noise radiated by fixed
isolated profiles in a wind tunnel, since they are much simpler to study than rotating blades, and
they may give valuable information for modeling inflow effect and fan noise in general. A good
review of the work carried out on the noise of fixed airfoils is presented by Blake (1986, §11-12).

69
Fan Noise Installation Effects

Photo courtesy of Howden Buffalo, Inc.


2.2.3 Reduction methods

Various means that allow for the reduction of fan noise associated with bad inflow conditions have
been presented for different types of fans in Section 1.3. These means may be used for minimizing
acoustic system effects. Some of them are detailed here.

The first option is to keep upstream obstacles and singularities away from the fan in order to
have a reasonably homogeneous and non turbulent flow at the impeller inlet. Obstacles very close
to the fan discharge must also be moved away. Figure 53 shows examples of good and bad fan
installations (Sharland 1988). It is recommended to keep a length of at least 1D between an elbow
and the fan inlet. This distance is a minimum, and it is advisable to increase this distance to 2 or
3D for axial flow fans.

As already mentioned in Section 2.2.1, some axial fans are more susceptible to inflow system
effects than others. It seems that axial impellers with broad chord blades, designed according to a
forced vortex law (mean load is higher at the blade tip), are less sensitive to turbulence ingestion
than narrow blades uniformly loaded along the span. The influence of the blade chord length on
noise system effects may be explained from the unsteady airfoil theory (Kemp and Sears 1955,
478-483).

70
Effect of Inflow Conditions

Incorrect Correct

Plain inlet Coned or bellmouth inlet

Slack flexible connector Taut flexible connector

Duct
diameter
minimum

Flow

Impeller immediately after bend Settling length before impeller

Figure 53 - Suggestions for Improving Fan Installation in Ductworks

The aerodynamic transfer function between the inlet flow velocity and the fluctuating forces on the
blades depends on the non-dimensional frequency ω = Cν/2U, where C is the chord length, ν is
the frequency of the upstream disturbance, and U is the relative mean flow velocity. The Sears
function, S(ω), which accounts for the response of the blade to the disturbance, is a complex
function whose amplitude decreases when the frequency ω increases (Figure 54). For a given

71
Fan Noise Installation Effects

disturbance frequency, v, the amplitude of the blade response, S(ω), is smaller when C is larger,
which may explain why large tip chord blades are less sensitive to turbulent inflow conditions than
narrow blades.

Figure 54 .5 IMAG.
Sears Function S(ω) 2
3 9 1 REAL
-.5 8 .5 1
10
7
4 5 Cv
6 .4 .4 =0
-.25 2U
.1
S (ω)

Bolton mentions several studies in which the authors try to minimize the blade response to inlet
turbulent flow by optimizing the fan design (1992, 77-88). A question remains about the
generalization of these results.

The reduction of the acoustic system effect due to bad inflow conditions can be obtained first by
the minimization of the upstream flow disturbances. A new design of the blades, such as increasing
the blade chord, may be attempted, but success is not guaranteed because of a lack of
understanding in this field.

2.3 Acoustic Loading Effect

2.3.1 Origins and main characteristics

Fan sound power levels depend on the acoustic impedance (see definition in Section 2.3.2) of both
the inlet and the outlet duct systems, i.e., on the way the sound waves emitted by the fan are
transmitted into the installation. This effect occurs mainly at low frequencies, such as below the
first cut-on frequency of the duct. For a circular duct, the cut-on frequency is f = 200/D, where D is
the duct diameter, in meters. As a result, with large ducts, this effect is comparatively less important
in the audible frequency range than on smaller systems.

This mechanism may be described as follows: the sound waves emitted by the fan are partly (or
sometimes totally) reflected by singularities like elbows, sudden changes in section, duct openings,
or obstacles in a way that depends on frequency. The reflected waves modify the sound power
radiated by the fan, both in the duct where the reflectors are, and in the duct at the opposite side.
This loading effect, which involves tonal and broadband noise, is characterized according to
frequency by amplification or attenuation of the fan power levels that are initially measured on a
standardized test rig. If, for instance, the frequency of a BPF tone coincides with an acoustic
resonance frequency of the duct, the amplitude of the tone may increase to more than 20 dB.
Conversely, this amplitude may strongly decrease if the BPF coincides with an anti-resonance of
the system.

72
Acoustic Loading Effect

2.3.2 Prediction method

2.3.2.1 Basis of the method

To deal with low-frequency sound wave propagation (or plane-wave propagation) in ducts and to
predict the influence of the system on the fan sound power level, an electrical analogy is
customarily used, in which the fan is equivalent to a source of current or voltage with an internal
impedance, and the system to an electrical circuit including components with their own electric
impedances. With this analogy, sound pressure, p, is equivalent to voltage, and volume velocity,
U = u.S (where u is the particle velocity and S is the duct section area), is equivalent to current.
The acoustic impedance, Z, in a section of the duct is Z = p/U. Like the electrical impedance, the
acoustic impedance is a complex variable that depends on frequency. It has a direct relation to the
sound reflection coefficient. The impedance of a duct system accounts for the sound reflection by
all the singularities in series or in parallel in the circuit.

Referring to the works of Baade (1976, 5-15), continued by Bolton and Margetts (1984), among
others, a circuit including a fan connected to an inlet and an outlet duct may be represented by
Figure 55. The fan sound power emitted in the outlet duct may be written as:

pv 2R2
W2 = 2
Equation 25
Z1 + ZV + Z2

Where:
W2 : outlet fan sound power (watts)
pv: fan internal pressure (or fan source strength)
Z1: acoustic impedance of the inlet duct
Z2: acoustic impedance of the outlet duct
Zv: acoustic impedance of the fan seen from the outlet duct
R2: real part of Z2

Figure 55
q1 q2
Acoustic Impedance Representation
of a Ducted Fan ZV Pv

Z1 Z2

73
Fan Noise Installation Effects

Equation 25 only applies in the plane-wave propagation frequency range, which is below the first
cut-on frequency defined in Section 2.3.1. The inlet fan sound power W1 is obtained by permuting
subscripts 1 and 2 in Equation 25. Fan internal pressure, pv, and fan impedance, Zv, do not depend
on the layout of the inlet and outlet ducts. They are functions of the fan operating point only.

Equation 25 clearly establishes that the outlet (or the inlet) fan power depends on both the inlet
and outlet duct geometry.

2.3.2.2 Description of the method

This section details the method of predicting the acoustic loading effect and presents some
measurement-prediction comparisons. In this presentation, we only consider the inlet sound power
while the outlet duct remains unchanged, or conversely, the outlet sound power with the inlet duct
unchanged. This greatly simplifies the implementation of the method. To treat the whole problem
and predict both the inlet and outlet sound levels, Abom, Boden, and Lavrentjev describe an
approach which may be adopted (1992, 359-364).

When the outlet sound power is considered for fixed inlet duct conditions, the equivalent electrical
diagram is shown in Figure 56, and Equation 25 is rewritten as:

p 2 v Re(Z2 )
W2 = 2
Equation 26
Z2 + Z v

Where Zv is the impedance of the fan + inlet duct, seen from the outlet side.

Similarly, the inlet sound power is obtained from Equation 26 by permuting subscripts 1 and 2 and
considering that Zv is the impedance of the fan connected to the outlet duct.

Figure 56
Impedance Representation of a -q q
One-Port Source Model Zv

p
Pv Z

74
Acoustic Loading Effect

The assessment of the terms in Equation 26 is made using a method described by Boden, Abom,
and Labrentjev (1992, 351-358). Ignoring subscripts 1 and 2, fan source strength, pv, is then:

⎛ Z ⎞
pv = p ⎜ 1 + v ⎟ Equation 27
⎝ Z ⎠

Where p is the acoustic pressure measured in an arbitrary reference section close to the fan.
Impedances Zv and Z have to be determined in the same reference section.

Measurement of the acoustic impedances can be done using a classical two-microphone


technique with an experimental set-up shown in the diagram of Figure 57. The component (fan or
duct singularity) whose impedance is to be measured is placed at one end of the duct, and a loud-
speaker is placed at the other end. Two microphones are mounted flush axially on the duct wall in
a section at the middle of the duct; the distance between the microphones is typically 5 cm. The
loudspeaker is supplied with a broadband random signal, and the complex transfer function
between the two microphone signals is measured using a dual-channel signal analyzer. The
acoustic impedance is linked to the transfer function by a simple relation (Guedel, Andre, and
Freynet 1998).

Figure 57
Experimental Set-up for the Determination
of Acoustic Impedances

Component to be tested

Loudspeaker

Microphones

It can be highlighted that the fan impedance, Zv, can be determined without the fan running since
experience shows that the reflection coefficient of the fan does not depend much on the impeller
rotation. This result has been checked on both an axial and a centrifugal fan (Guedel, Andre, and
Freynet 1998), as well as by other authors. In the same way, the duct impedance is nearly
independent of the flow velocity when the Mach number is moderate; specifically, below 0.3. It is
therefore acceptable to determine the acoustic impedance with no impeller rotation and no flow,
which considerably simplifies the experimental process.

The acoustic impedance of a duct system may also be obtained by calculation, using the transfer
matrix technique, since the circuit may be modeled by a number of singularities connected to each

75
Fan Noise Installation Effects

other by straight ducts. The sound reflection coefficients of the singularities have to be known in
this case.

The fan sound power level in each frequency band can be deduced from Equation 26:

⎛ Re(Z ) ⎞
Lw = Lpv + 10 lg ⎜ ⎟ Equation 28
⎜ Z +Z2 ⎟
⎝ v ⎠

Where Lpv is the source pressure level, which is obtained from Equation 27 and rewritten as:

⎛ Z ⎞
Lpv = Lp + 10 lg ⎜ 1 + v ⎟
⎝ Z ⎠

Lp is the sound pressure level measured in the reference section with the fan in operation.

The acoustic loading effect, which is the difference in each frequency band between the sound
power levels of the fan in the system and in the standardized installation at the same operating
point, is derived from Equation 28:

⎛ Re(Z ) ⎞ ⎛ Re(Z ) ⎞
dLw = 10 lg ⎜ ⎟ − 10 lg ⎜ ref
⎟ Equation 29
⎜ Z+Z ⎟ 2
⎜ Z +Z 2 ⎟
⎝ v ⎠ ⎝ ref V ⎠

Where Zref is the impedance of the standardized installation. Equation 29 accounts for the fact that
Zv and Lpv are independent of the duct configuration, Lpv varying only with fan speed and operating
point.

76
Acoustic Loading Effect

15

10 Experiment
Prediction
5
dLw (dB)

a) 0D 0
100 125 160 200 250 315 400 500
-5

-10

-15
f (Hz)
15

10 Experiment
Prediction
5
dLw (dB)

b) 0.5D 0
100 125 160 200 250 315 400 500
-5

-10

-15
f (Hz)
15

10 Experiment
Prediction
5
dLw (dB)

c) 3D 0
100 125 160 200 250 315 400 500
-5

-10

-15
f (Hz)

Figure 58 - Comparison of Measured and Predicted Acoustic Effect on a Centrifugal


Fan with a 90° Elbow at Various Distances from the Fan Outlet

77
Fan Noise Installation Effects

2.3.2.3 Comparison of predicted and measured system effects

Figure 58 is a comparison of the measured and predicted acoustic system effects on a centrifugal
fan connected to a bent duct of diameter D = 350 mm on its outlet side (Guedel, Andre, and
Freynet 1998). The distance from the fan exit to the 90° elbow varies from 0 to 3D. Since the elbow
is at the fan discharge, the system effect associated with inflow conditions is negligible compared
to the acoustic impedance effect.

Measured and predicted curves show quite similar trends, even if the prediction sometimes
magnifies the effect. This result and other comparisons not shown here prove that the prediction
model may be considered validated as the maximum uncertainty of the result in one-third octave
bands being estimated at 2 to 3 dB. This is satisfactory if we consider values of overall system
effects that may reach 20 dB or more in some cases.

2.3.3 Reduction methods

Means of reduction in the acoustic loading effect result from the prediction method described
above. Unlike system effects, due to inflow conditions, increasing the distance between the fan and
the reflecting singularity does not guarantee a reduction of the system effect unless the reflector is
moved far enough. In this case, the reflected waves reaching the fan are strongly attenuated and
do not modify the fan sound power like in a standardized test duct equipped with an anechoic
termination.

The prediction method described above allows for optimization of the duct system in order to
minimize the loading effect, especially at the BPF harmonics if the amplitude is high.

2.4 Concluding remarks

Acoustic system effects may significantly increase the fan sound power level that is measured on
a standardized test installation. This effect has two origins: one due to the perturbation of the flow
field at the fan inlet by the system, and the other due to the sound wave reflections by the
singularities of the circuit.

The effect associated with the inflow conditions is difficult to predict since it is related to the fan
noise generation mechanisms which are relatively well known, but cannot be predicted accurately
at this time. Means to reduce this effect are nevertheless known; they consist of moving the
obstacles away from the fan in order to minimize the inflow disturbance.

The acoustic loading effect by the system is easier to predict from a calculation of the acoustic
impedance coupling between the fan and the duct system. The prediction method presented here
is simple to implement, especially because the acoustic impedances may be evaluated with no
flow and compare favorably with experiment. Some discrepancies still exist between the predicted
and measured results in the case of strongly marked acoustic resonances, but it is not a drawback,
as the model provides the correct trend.

78
Annex A. Fan Terminology

Geometry

camber angle: angle between the tangents to the blade at the leading edge and at the trailing
edge (Figure A.1).

casing: shroud of an axial-flow fan or volute of a centrifugal fan.

chord: segment between the leading edge and the trailing edge (Figure A.1).

impeller: rotating part of the fan.

leading edge: upstream part of the blade section (Figure A.1).

pitch angle: angle between the chord line and the rotation plane for axial fans and between the
chord line and the tip radius for centrifugal fans.

pressure side: concave side of the blade.

solidity: ratio of the chord length over the blade spacing.

stator: fixed part of the fan, such as guide vanes on axial fans or volute on centrifugal fans.

suction side: convex side of the blade.

tip clearance: radial gap between the blade tip and the casing on an axial-flow fan.

trailing edge: downstream part of the blade section (Figure A.1).

twist: a twisted blade has its pitch angle that varies along the span.

Performance

angle of attack: angle between the relative inlet velocity w1 and the tangent to the blade leading
edge (Figure A.2).

79
Fan Terminology

fan characteristic (or fan curve): curve relating fan pressure to volume flow.

fan efficiency: ratio of the fan power over the fan shaft power. Fan static efficiency is the ratio of
the fan static power over the fan shaft power.

fan outlet dynamic pressure: pressure defined as Pd2 = 0.5ρV22, where V2 = Qv / S2 (Qv = volume
flow, S2 = area of the fan outlet section).

fan power: product of the volume flow by the fan pressure at a given operating point.

fan shaft power: mechanical power supplied to the fan impeller.

fan static power: product of the volume flow by the fan static pressure.

fan static pressure: by convention, difference between the fan total pressure and the fan outlet
dynamic pressure ΔPs = ΔPt - Pd2.

fan total pressure (or fan pressure): difference between the total pressure at the fan outlet and
the total pressure at the fan inlet ΔPt = Pt2 - Pt1.

motor slip: fan rotational speed may vary with the operating point because of the motor slip. The
speed decreases when the resisting torque, i.e. the fan power, increases.

operating point: point on the fan characteristic.

velocity triangle: gives the magnitude and the direction of the flow velocity across the impeller.
Figure A.2 shows an example of velocity triangles at the leading edge and trailing edge of an axial
impeller (Daly 1992), in which:

v is the absolute velocity (flow velocity in the fixed coordinate system)


u is the blade velocity at radius considered
w is the relative velocity (flow velocity in the coordinate system rotating with the blades)

80
Fan Terminology

Figure A.1 Trailing edge


Blade Section
Camber angle

rd
ho
C

Leading edge

Figure A.2 -u 2
Velocity Triangles on an
2
Axial Fan

S
W2 V2
(Daly 1992)

S1
V1
W1

-u 1

81
Annex B. Acoustics Terminology

A-weighted sound levels: levels obtained after applying A-weighting corrections to take into
account the human ear response to noise. Figure B.1 shows the A-weighting curve that applies to
octave band or one-third octave band levels. The A-weighted overall sound level (dBA) is the
overall sound level calculated from A-weighted levels over the spectrum.

acoustic impedance: ratio of the sound pressure, p, over the particle velocity, u, at the same
point.

airborne and structure-borne noise: a fan radiates airborne noise from its openings and casing.
Structure-borne noise is due to vibrations of the fan supports and ducts, which emit noise (Figure
B.2).

background noise: noise remaining when the sound source to be tested is stopped. A difference
of at least 6 dB is required between the source level and the background noise level to consider
the measurement as valid.

blade passage frequency: a fan with evenly spaced blades radiates tonal noise at the
fundamental and the first harmonics of the blade passage frequency BPF = B × N (where B =
number of blades, N = rotational speed in Hz).

overall sound level: level obtained by adding up the levels of all the frequency bands of a
spectrum. The summation is logarithmic: LA+B =10 lg(100.1L +100.1L ).
A B

particle velocity (or acoustic velocity): velocity associated with the motion of the air molecules
around the sound wave front. This velocity, usually called u, must not be mistaken for the speed
of sound, c0, which is the velocity of propagation of the wave front (c0 = 343 m/s for air at 20°C).

sound intensity: sound energy that flows through a unit area in a unit time (watts/m2). Sound
intensity is a vector in the direction of the sound wave propagation. For a free progressive sound
wave in which u and p are in phase, the average intensity in the direction of the wave propagation
is I = <pu>, where the brackets stand for the time average of the product. For a free-traveling plane
sound wave, the intensity is I = p2 / ρc0, in which ρ is the air density and c0 is the sound speed.
The sound power, W, may be obtained from the sound intensity from the following formula:

82
Acoustics Terminology

W = ∫ IndS
S

Where S is a surface enveloping the sound source and In is the component of the intensity vector
normal to S in each point.

⎛ W ⎞
sound power level: defined by LW = 10 lg ⎜ −12 ⎟
⎝ 10 ⎠
Where W is the sound power in watts. Unlike the sound pressure level, the sound power level does
not depend on the measurement distance, the source directivity or the acoustic properties of the
site. It accounts for the sound energy radiated by the source in the whole space, whatever the
source location is in the test site. The sound power level is also in dB, so that it is essential to
mention whether a sound level in dB refers to a pressure level Lp or a power level Lw.

2
⎛ p ⎞
sound pressure level: defined by: Lp = 10 lg ⎜ −5 ⎟
⎝ 2 × 10 ⎠
Where p is the sound pressure in Pa. The sound pressure level Lp in dB depends on parameters
like the distance between the noise source and the measurement point, the source directivity and
the reverberating properties of the test site.

spectrum: graph of the sound pressure or sound power level as a function of frequency. Three
types of spectra are used in acoustics: narrow-band spectrum representing the sound energy
distribution in constant bandwidths ranging from a fraction of Hz to several Hz (power spectral
density is the spectrum for a bandwidth of 1 Hz), octave band and one-third octave band
spectra, in which the bandwidth is non constant but varies proportionally to frequency. Figure B.3
compares the narrow-band and the one-third octave band spectra measured on the same fan.

wavelength: distance between two maximums (or minimums) of a sound wave at a given
frequency. Wavelength λ is linked to frequency f by λ = c0 / f.

83
Acoustics Terminology

5
0
A-weighting (dB)

-5
-10
-15
-20

-25
-30
50

100

200

400

800

1600

3150

6300

12500
f (Hz)
Figure B.1 - A-Weighting Curve

Figure B.2
Airborne and Structure Borne Noise of a Fan
Airborne noise:
noise radiated by the fan openings

.
.
. .
. .
.

Structure borne noise:


noise due to the vibrations of fan supports

84
Acoustics Terminology

70 Narrow-band spectrum
65 One-third octave band spectrum
60
55
Lp (dB)

50
45
40
35
30
100 1000 10000
f (Hz)
Figure B.3 - Comparison of the Narrow-Band and the
One-Third Octave Band Spectra of the Same Noise

85
Annex C. Forward-Curved Centrifugal Fan Design

Design characteristics of impeller and volute of FC centrifugal fans are presented here. These
values of the design parameters, which result from experiments carried out at CETIAT or
elsewhere, guarantee satisfactory performance and low-noise level. The symbols used in the
following tables are defined in the list of symbols and Figures C.1 and C.2 below.

D1 / D2 = 0.8 to 0.9

D / D1 = 1

W / D2 = 0.3 to 0.7 (beyond 0.7, sound level increases)

C / D2 = 0.08 to 0.14

C / t = 0.9 to 1.4 (0.9 at low flow and 1.4 at high flow)

γ = 24 to 30°

δ = 70 to 90°

Wv / W = 1.1 to 1.4

j ≥ 0.1 (preferably 0.12 to 0.15 when low noise level is required)

θ0 = 60 to 90° (60° at low flow)

Is = 0.1 to 0.12 (0.1 at low flow)

Σr / Σv = 2 to 4 (3.5 to 4 at low flow)

86
Forward-Curved Centrifugal Fan Design

List of symbols:

D1: inner impeller diameter Wv: volute width (single-width impeller)


D2: outer impeller diameter j: non dimensional cutoff clearance (j = r0/D2 - ½)
D: bellmouth exit diameter r0: volute radius at cutoff
W: impeller width (single-width impeller) θ0: cutoff angle (Figure C.2)
C: blade chord length Is: volute expansion index defined by:
t: blade spacing r(θ) = r0eIsθ (where θ is in radians)
γ: blade pitch angle Σr: outer impeller surface area (= πD2W)
δ: camber angle
Σv: fan outlet area

D2 β1
Figure C.1 D1
Impeller Design
β2
t

Figure C.2
Volute Design

r (θ)
θ0
r0

87
References

Abom, M., H. Boden, J. Lavrentjev. 1992. Source Characterization of Fans Using Acoustic 2-Port
Models. Proceedings of the Fan Noise Symposium, CETIM Senlis 359-364.

AICVF. 1997. Bruit des Équipements. Collection des Guides de l’AICVF. Editions PYC Livres.

Air Movement and Control Association. 1990. Standard 301-90: Methods for Calculating Fan
Sound Ratings from Laboratory Test Data.

Air Movement and Control Association. 1990. Publication 201-90: Fans and Systems.

Air Movement and Control Association. 1996. Standard 300-96: Reverberant Room Method for
Sound Testing of Fans.

Ameziane, 1992. Contribution à la Prédiction et la Réduction du Bruit Aérodynamique d’une


Turbomachine de Type Hélico-Centrifuge. Thése de Doctorat Université Paul Sabatier,
Toulouse.

Ameziane, H., S. Pauzin, J. P. Guilhot, C. Biben. 1992. Noise Reduction in a Mixed-Flow Fan.
Proceedings of the Fan Noise Symposium, CETIM Senlis 179-186.

Amiet, R. K., R. W. Patterson. 1979. Noise of a Model Helicopter Rotor Due to Ingestion of
Turbulence. NASA Report 3213.

Baade, P. K. 1976. Effects of Acoustic Loading on Axial Flow Fan Noise Generation. Noise Control
Engineering 8(1): 5-15.

Besombes, M. 1993. Validation du Principe du Contrôle Actif à la Source Appliqué à Une


Turbomachine Centrifuge de Conception Industrielle. Rapport D’étude CETIM n° 172970.

Blake, W. K. 1986. Mechanics of Flow-Induced Sound and Vibration, Volume II. Chapters 11 and
12. Academic Press, Inc.

Boden, H., M. Abom, J. Lavrentjev. 1992. Experimental Methods for Determining the Source
Characteristics of Fans Modelled as Acoustic One-Port Sources. Proceedings of the Fan Noise
Symposium, CETIM Senlis 351-58.

Bolton, A. N., E. J. Margetts. 1984. The Influence of Impedance on Fan Sound Power. IMechE
Conference on Installation Effects in Ducted Fan Systems, London.

88
References

Bolton, A. N. 1992. Fan Noise Installation Effects. Proceedings of the Fan Noise Symposium,
CETIM Senlis 77-88.

Bommes, L. 1992. A Nomogram for Use in the Assessment of Fan Noise. HLB 43:598-604.

Bommes, L., R. Grundmann, k. Klaes, C. Kramer. 1995. Effects of Blade Design on Centrifugal Fan
Noise and Performance. Noise Control Engineering Journal 43(4): 91-101.

Bridelance, J. P. 1982. Etude Aéroacoustique des Spectres de Raies Générés par les Ventilateurs
Axiaux en Régime Subsonique. Thése de Docteur-Ingénieur ENSAM.

British Standard 848: Part 2. 1985. Fans for General Purposes. Methods of Noise Testing.

Burgain, T. 1998. Contribution du Bruit Propre des Pales au Bruit à Large Bande des Ventilateurs
Centrifuges. Thèse de Doctorat ECL, Lyon.

Carolus, T. 1992. Acoustic Performance of Low Pressure Axial Fan Rotors with Different Blade
Chord Length and Radial Load Distribution. Proceedings of DGLR/AIAA 14th Aeroacoustics
Conference 809-15.

Cory, W. T. W. 1992. Acoustic Similarity Laws for the Prediction of Industrial Fan Sound Levels.
Proceedings of the Fan Noise Symposium, CETIM Senlis: 305-28.

Croba, D., J. L. Kueny. 1992. Unsteady Flow Computation in a Centrifugal Pump: Coupling of the
Impeller and the Volute. Proceedings of the Fan Noise Symposium, CETIM Senlis 221-28.

Curle, N. 1955. The Influence of Solid Boundaries Upon Aerodynamic Sound. Proceedings of the
Royal Society A231:505-14.

Daly, B. B. 1992. Wood’s Practical Guide to Fan Engineering. Woods of Colchester Limited.

DIN 45 635 Part 38. 1986. Measurement of Noise Emitted by Machines. Enveloping Surface
Method, Reverberation Room Method and In-Duct Method. Fans.

Eck, B. 1973. Fans. Pergamon Press.

Ffowcs Williams, J. E., D.L. Hawkings. 1969. Sound Generation by Turbulence and Surfaces in
Arbitrary Motion. Philosophical Transactions of the Royal Society A264.

Fournier, F. 1988. Mise au point d’une méthode de calcul adaptée au bruit des fenestrons
d’hélicoptéres. Thése de Doctorat ECL 88-09, Lyon.

Fuest, T., T. Carolus. Comparative Representation of Measured Boundary Layer Fluctuating


Pressures and Sound Power of Two Different Axial Fans. Proceedings of Inter-Noise 95:97-100.

89
References

Fukano, T., Y. Kodama, Y. Senoo. 1977. Noise Generated by Low Pressure Axial Flow Fans, Part
1. Journal of Sound Vibration 50:63-74.

Fukano, T., Y. Kodama, Y. Takamatsu. 1978. Noise Generated by Low Pressure Axial Flow Fans,
Part 3. Journal of Sound and Vibration 56:261-77.

Fukano, T., Y. Takamatsu, Y. Kodoma. 1986. The Effects of Tip Clearance on the Noise of Low
Pressure Axial and Mixed Flow Fans. Journal of Sound and Vibration 105:291-308.

Fukano, T., Y. Kodama. 1992. Prediction of Sound Power of Low Pressure Axial and Diagonal Flow
Fans. Proceedings of the Fan Noise Symposium, CETIM Senlis 105-112.

Gautier, D., J. F. Combes, A. Laporta, A. Audant. 1998. Le Ventilateur Tangentiel: Le Nouveau


Secret Minceur des Climatiseurs. Revue Epure n°57. Direction des Etudes et Recherches
d’EDF.

Graham, J. Barrie. 1992. The Status of Fan Noise Measurement in North America. Proceedings of
the Fan Noise Symposium, CETIM Senlis:293-300.

Gray, A. J. 1994. Intercomparison of Open-Inlet / Open-Outlet Measurements on Fans, Part 3.


Community Bureau of Reference Report EEC009.

Guedel, Alain. 1989. Bruit des Ventilateurs Hélicoïdes. Synthése bibliographique. Document
Technique CETIAT VTL 2.

Guedel, Alain, F. Mairet. 1990. Bruit des Moteurs électriques de faible puissance utilises sur les
ventilo-convecteurs. Rapport d’étude CETIAT 890444/4.

Guedel, Alain, N. Yazigi. 1992. Influence of Blade Sweep on the Noise of a Low-Speed Tubeaxial
Fan. Prodeedings of the Fan Noise Symposium, CETIM Senlis 167-78.

Guedel, Alain. 1993. Influence de la Géométrie d’un Ventilateur Hélicoïde Sur Son Niveau de Bruit.
Note Technique CETIAT NTV 93078.

Guedel, Alain. 1995a. Bruit des Ventilateurs Axiaux couplés à un Échangeur de Chaleur.
Acoustique et Techniques n° 3 et. 4.

Guedel, Alain. 1995b. Bruit des Ventilateurs Centrifuges Cage D’écureuil. Note Technique CETIAT
NT 95098.

Guedel, A., M. Freynet. 1996. Optimisation Aéroacoustique des Ventilateurs Tangentiels: Étude
Paramétrique Sur Une Roue de Diamétre 100 mm. Note Technique CETIAT NT 96001.

Guedel, A., M. Freynet, L. Boiteux. Optimisation Aéroacoustique des Ventilateurs Tangentiels:


Étude Paramétrique Sur Une Roue de Diamétre 60 mm. Note Technique CETIAT NT 96059.

90
References

Guedel, A., M. Freynet. 1997a. Normalisation Acoustique des Ventilateurs: Comparison des
Niveaux de Puissance Acoustique en Champe Libre mesurés Suivant les Normes DIN, BS, et.
NF. Note Technique CETIAT. NTV 97195.

Guedel, A., M. Freynet. 1997b. Réduction du Bruit Dans des Ensembles Compacts Échangeur-
Ventilateur. Note Technique CETIAT NT 97289.

Guedel, A. 1997. Effets D’installation sur des Ventilateurs Hélicoïdes couplés à un échangeur de
Chaleur. Note Technique CETIAT NT 97151.

Guedel, A., G. Perrin, M. Freynet. 1997. Réduction du Bruit des Ventilateurs Hélicoïdes de
Climatiseurs. Note Technique CETIAT NT 97 232.

Guedel, A., C. Andre, M. Freynet. 1998. Effets D’installation des Ventilateurs: Influence D’un
Coude en Entrée ou en Sortie de Ventilateurs. Note Technique CETIAT NT 97151.

Hofe, R. V., G. E. Thien 1987. Low Noise, Compact and Efficient Cooling System With Tangential
Flow Fan. SAE Paper 870983 3.80-3.89.

Hofe, R. V., G. E. Thien. 1992. Quiet and Efficient Tangential Flow Fan in Compact Heat Exchanger
Application. Proceedings of the Fan Noise Symposium, CETIM Senlis 237-244.

Holste, F., W. Neise. 1992. Experimental Comparison of Standardised Sound Power Measurement
Procedures for Fans. Journal of Sound Vibration 152:1-26.

Horlock, J. H. 1968. Fluctuating Lift Forces on Aerofoils Moving Through Transverse and
Chordwise Gusts. ASME Journal of Basic Engineering 90:494-500.

Howe, M.S. 1978. A Review of the Theory of Trailing Edge Noise. Journal of Sound and Vibration
61:437-465.

Hunnaball, P. J. 1992. Control of Tonal Noise Generation in Axial Flow Fans by Optimising
Geometry of Fixed and Rotating Components. Proceedings of the Fan Noise Symposium,
CETIM Senlis 475-82.

ISO 9614-1. 1993. Determination of Sound Power Levels of Noise Sources Using Sound Intensity,
Part 1: Measurement at Discrete Points.

ISO 9614-2. 1993. Determination of Sound Power Levels of Noise Sources Using Sound Intensity,
Part 2: Measurements by Scanning.

ISO 3744. 1994. Acoustique - Détermination des niveaus de puissance acoustique émis par les
sources de bruit - Méthodes d’expertise pour les conditions de champ libre au-dessus d’un plan
réfléchissant.

91
References

ISO 10302. 1996. Acoustics - Method for the Measurement of Air-Borne Noise Emitted by Small
Air-Moving Devices.

ISO 5136. 1999. Acoustics - Determination of Sound Power Radiated Into a Duct by Fans and
Other Air-Moving Devices - In-Duct Method.

ISO Technical Committee 117. 1996. Ventilateurs Industriels - Terminologie. Projet de Norme
Internationale ISO/DIS 13349.2

Kaji, S., T. Okazaki. 1970. Axial-Flow Compressor Noise Studies. Journal of Sound Vibration
13(3):281-307.

Kemp, N. H., W. R. Sears. 1955. The Unsteady Forces Due to Viscous Wakes in Tubomachines.
Journal of Aeronautical Science 22:478-83.

Konieczny, P., Bolton, J.S. 1995. Design of Low-Noise Centrifugal Blowers. Noise Control
Engineering Journal 43(4) 103-27.

Leze, F., M. Besombes, J. Tourret, M. Pluviose, M. Bertinier. 1992. Visualization of the Dynamic
Pressure Field in a Centrifugal Fan in the Study of Noise Generation Mechanisms. Proceedings
of the Fan Noise Symposium, CETIM Senlis 213-220.

Lighthill, M. J. 1952. On Sound Generated Aerodynamically; General Theory. Proceedings of the


Royal Society A211:564-87.

Lowson, M. V. 1970. Theoretical Analysis of Compressor Noise. Journal of the Acoustical Society
of America 47:371-85.

Lowson, M. V., J. B. Ollerhead. 1969. A Theoretical Study of Helicopter Rotor Noise. Journal of
Sound Vibration 69:197-222.

Madison, R. D. 1949. Fan Engineering 5th ed. Buffalo Forge Company, Buffalo, New York.

Marteel, S., B. Desmet, H. Wullens. 1992. Aeroacoustic Study of the Line Spectrum Generated by
a Subsonic Axial Flow Fan. Proceedings of the Fan Noise Symposium, CETIM Senlis 141-148.

Mongeau, L., D. E. Thompson, D. K. McLaughlin. 1995. Method for Characterising Aerodynamic


Sound Sources in Turbomachines. Journal of Sound Vibration Vol. 181(3):369-389.

Morinushi, K. 1987. The Influence of Geometric Parameters on FC Centrifugal Fan Noise. Journal
of Vibration, Acoustics, Stress and Reliability in Design 109:227-34.

Neise, W. 1982. Review of Noise Reduction Methods for Centrifugal Fans. Trans ASME Journal of
Engineering for Industry 104:151-161.

Neise, W. 1988. Fan Noise - Generation Mechanisms and Control Methods. Proceedings of Inter-
Noise 88:767-76.

92
References

Neise, W. 1989. Noise Rating of Fans on the Basis of the Specific Sound Power Level. Tenth
Australian Fluid Mechanics Conference, Melbourne.

Neise, W., G. H. Koopmann. 1988. Active Source Cancellation of Blade Tone Fundamental and
Harmonics in Centrifugal Fans. Journal of Sound and Vibration 126(2): 209-20.

Neise, W. 1992. Review of Fan Noise Generation Mechanisms and Control Methods. Proceedings
of the Fan Noise Symposium, CETIM Senlis 45-56.

NF S31-021. 1982. Acoustique - Mesurage en Plateforme du Bruit émis par les Ventilateurs a
Enveloppe. Méthode du Caisson Réduit au Refoulement.

Perrin, G. 1997. Modélisation des Ventilateurs Type Cage D’écureuil. Note Technique CETIAT NT
97006.

Ponsonnet, P. 1972. Bruit des ventilateurs et calcul acoustique des installations aérauliques. Ed.
Solyvent-Ventec.

Raffaitin, C. 1995. Caractérisation des Sources de Bruit Aérodynamique sur un Ventilateur


Centrifuge à flux axial. Thése de Doctorat ECL, Lyon.

Reznicek, M. E., L. Mongeau. 1995. Method for Characterising Aerodynamic Sources of Sound
with Applications to Computer Cooling Fans. Proceedings of Inter-Noise 95:265-8.

Riera-Ubiergo, J. 1988. Méthodes Pratiques de Réduction du Bruit des Ventilateurs Centrifuges:


Analyse Bibliographique. Document Technique CETIAT VTL 1.

Sharland, I. J. 1964. Source of Noise in Axial Flow Fans. Journal of Sound Vibration 1:302-322.

Sharland, I. 1988. Wood’s Practical Guide to Noise Control. Published by Woods Acoustics.

Szentmartony, T., I. Kurutz, G. Kosco. 1992. Success, Failure and Possibility in Silencing Radial
Flow Fans. Proceedings of the Fan Noise Symposium, CETIM Senlis 271-78.

Thompson, M. C., K. Hourigan, A. N. Stokes. 1992. Prediction of the Noise Generation in a


Centrifugal Fan by Solution of the Acoustic Wave Equation. Proceedings of the Fan Noise
Symposium, CETIM Senlis 197-204.

Tournoy, D., E. Bruyere. 1992. Reduction of the Noise Generated by Industrial Centrifugal Fans:
Use of a Quarter Wavelength Resonator. Proceedings of the Fan Noise Symposium, CETIM
Senlis 279-85.

Tyler, J. M., T.G. Sofrin. 1962. Axial Flow Compressor Noise Studies. SAE Trans. 70.

Weidemann, J. 1971. Analysis of the Relations Between Acoustic and Aerodynamic Parameters
for a Series of Dimensionally Similar Centrifugal Fan Rotors. NASA Technical Translation TT F-
13. 798.

93
AIR MOVEMENT AND CONTROL
ASSOCIATION INTERNATIONAL, INC.
30 West University Drive
Arlington Heights, IL 60004-1893 U.S.A.
Tel: (847) 394-0150 Fax: (847) 253-0088
E-Mail : info@amca.org Web: www.amca.org

The Air Movement and control Association International, Inc. is a not-for-profit international association of the
world’s manufacturers of related air system equipment primarily, but limited to: fans, louvers, dampers, air
curtains, airflow measurement stations, acoustic attenuators, and other air system components for the industrial,
commercial and residential markets.

Você também pode gostar