Download as pdf or txt
Download as pdf or txt
You are on page 1of 74

Sound transmission properties of honeycomb panels and

double-walled structures

SATHISH KUMAR

Doctoral Thesis
Stockholm June 2012

Vinnova Centre of Excellence for ECO2 Vehicle Design

The Marcus Wallenberg Laboratory for Sound and Vibration Research


Department of Aeronautical and Vehicle Engineering

Postal Address Visiting Address Contact


The Royal Institute of Technology Teknikringen 8 Tel: +46 8790 9375
MWL/AVE Stockholm Email: sathish@kth.se
SE-100 44 Stockholm
Sweden
Akademisk avhandling som med tillstånd av Kungliga Tekniska Högskolan i Stock-
holm framläggs till offentlig granskning för avläggande av teknologie doktortexamen
torsdagen den 14 June 2012, 13:15 i sal F3 Lindstedtsvägen 26, KTH, Stockholm.

TRITA-AVE-2012:20
ISSN-1651-7660
ISBN 978-91-7501-334-3


c Sathish Kumar, June 2012
Tryck:US-AB
Abstract

Sandwich panels with aluminium face sheets and honeycomb core material have
certain advantages over panels made of wood. Some of the advantages of these con-
structions are low weight, good moisture properties, fire resistance and high stiffness-
to-weight ratio etc. As product development is carried out in a fast pace today, there is
a strong need for validated prediction tools to assist during early design stages. In this
thesis, tools are developed for predicting the sound transmission through honeycomb
panels, typical for inner floors in trains and later through double-walled structures
typical for rail-vehicles, aircrafts and ships.
The sandwich theory for wave propagation and standard orthotropic plate theory
is used to predict the sound transmission loss of honeycomb panels. Honeycomb is an
anisotropic material which when used as a core in a sandwich panel, results in a panel
with anisotropic properties. In this thesis, honeycomb panels are treated as being or-
thotropic and the wavenumbers are calculated for the two principal directions. The
wavenumbers are then used to calculate the sound transmission using standard or-
thotropic theory. These predictions are validated with results from sound transmission
measurements. The influence of constrained layer damping treatments on the sound
transmission loss of these panels is investigated. Results show that, after the damping
treatment, the sound transmission loss of an acoustically bad panel and a normal panel
are very similar.
Further, sound transmission through a double-leaf partition based on a honeycomb
panel with periodic stiffeners is investigated. The structural response of the periodic
structure due to a harmonic excitation is expressed in terms of a series of space harmon-
ics and virtual work theory is applied to calculate the sound transmission. The original
model is refined to include sound absorption in the cavity and to account for the or-
thotropic property of the honeycomb panels. Since the solution of the space harmonic
analysis is obtained in a series form, a sufficient number of terms has to be included
in the calculation to ensure small errors. Computational accuracy needs to be balanced
with computational cost as calculation times increases with the number of terms. A
new criterion is introduced which reduces the computational time by up to a factor
ten for the panels studied. For all the double-leaf systems analysed, the sound trans-
mission loss predictions from the periodic model with the space harmonic expansion
method are shown to compare well with laboratory measurements.

vii
Dissertation

The work presented in this Doctoral thesis was carried out within The Centre for ECO2
Vehicle Design at the Department of Aeronautical and Vehicle Engineering, The Royal
Institute of Technology (KTH) in Stockholm, Sweden.

This thesis consists of two parts. The first part gives an overview of the research with
a summary of the performed work. The second part collects the following scientific
articles:

Paper A. Leping Feng and Sathish Kumar, On Application of radiation loss factor in
the prediction of sound transmission loss of a honeycomb panel. International Journal of
Acoustics and Vibration, 17 (2012) 47-51.

Paper B. Sathish Kumar, Leping Feng and Ulf Orrenius, Predicting the sound trans-
mission loss of honeycomb panels using the wave propagation approach. Acta Acustica
united with Acustica, 97 (2011) 869-876.

Paper C. Sathish Kumar, Leping Feng and Ulf Orrenius, Modelling the sound trans-
mission through rib-stiffened double-leaf partitions with cavity absorption. Submitted
to Journal of Sound and Vibration, January 2012.

Paper D. Sathish Kumar, Leping Feng and Ulf Orrenius, Modelling the sound trans-
mission through rib-stiffened sandwich double-leaf partitions using space harmonic
analysis. Submitted to Journal of Sound and Vibration, May 2012.

ix
Contents of this thesis has been presented in the following
conferences

Parts of Paper A has been presented at the following conference:

• The 17th International Congress on Sound and Vibration, Cairo, Egypt, July 2010.

Parts of Paper B has been presented at the following conference:

• The 37th International Congress and Exposition on Noise Control Engineering,


Shanghai, China, October 2008.

Parts of Paper C has been presented at the following conference:

• Noise and Vibration: Emerging Methods, NOVEM- 2012, Sorrento, Italy, April
2012.

Division of work between the authors

Paper A. Sathish performed the measurements and provided data for computations,
Leping performed the computations and wrote the paper.

Paper B. Sathish performed the measurements, computations and wrote the paper
with inputs from Leping and Ulf.

Paper C. Sathish developed the model, performed the computations and wrote the
paper with inputs from Leping and Ulf.

Paper D. Sathish developed the model, performed the measurements, computations


and wrote the paper with inputs from Leping and Ulf.

x
Acknowledgments

The work presented in this thesis has been carried out within the Centre for ECO2 Vehi-
cle Design at the Department of Aeronautical and Vehicle Engineering, KTH, Sweden.
This thesis is a part of the research project “Coupling shape-vehicle body structure”
within the centre. The financial support provided by Vinnova, KTH and the industrial
partners is gratefully acknowledged.

To begin with, I would like to thank my main supervisor, Leping Feng for accepting
me as a PhD student and for his guidance, support and comments since the start of this
thesis. In addition many thanks to my co-supervisor Ulf Orrenius for his support, com-
ments on this work and for proofreading the manuscript of the papers and this thesis.
I would like to thank Per Wennhage for the useful discussions during the earlier stages
of this thesis. I would also like to thank Gunnar Ziwes from EURO-COMPOSITES R

S.A. for providing the honeycomb panels investigated in this thesis.


The measurements in this thesis wouldn’t be possible without the help of Kent
Lindgren and Danilo Prelevic. Computer support from Urmas Ross is also gratefully
acknowledged.
I also wish to express my gratitude to all my colleagues at the department and
within the ECO2 centre for providing a great working environment during the last 5
years. Thanks, in particular to Ciarán, Dmitry, Hans, Hao, Heiki, Jia, Tomas and Tristan.
Thank you guys for the useful and sometimes not-so-useful lunch time discussions.
Special thanks to Chenyang for introducing me to LYX and making my life easier while
writing this thesis.
To my family who has always stood by and supported me for all these years won-
dering when I would finish “school”, I want to thank you. I would also like to thank
my friends in Stockholm for making my stay pleasant for all these years.
But first and foremost, I would like to thank my wonderful wife Sharenya whose
love and moral support made this thesis possible. Thank you Kutty!

Sathish Kumar

Stockholm, June 2012

xi
Contents

I Overview and Summary 1

1 Introduction 3

2 Floor constructions in trains 7


2.1 Floating floors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Sandwich constructions and the sandwich effect . . . . . . . . . . . . . . 8
2.3 Honeycomb panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 Damping treatments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 Sound transmission through honeycomb panels 11


3.1 Wave propagation in sandwich structures . . . . . . . . . . . . . . . . . . 11
3.2 Sound transmission through orthotropic plates . . . . . . . . . . . . . . . 13
3.3 Radiation loss factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

4 Sound transmission through double-leaf partitions with periodic stiffeners 17


4.1 Space harmonic analysis - Isotropic panels . . . . . . . . . . . . . . . . . . 18
4.2 Space harmonic analysis - Sandwich panels . . . . . . . . . . . . . . . . . 22
4.3 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.3.1 Traditional convergence criterion . . . . . . . . . . . . . . . . . . . 25
4.3.2 New convergence criterion . . . . . . . . . . . . . . . . . . . . . . 26

5 Measurements 29
5.1 Loss factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2 Sound reduction index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.3 Vibration velocity level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5.4 Bending stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

6 Results and discussion 33


6.1 Honeycomb panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.2 Double-leaf partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

xiii
7 Conclusions and future work 51

8 Summary of appended papers 53


8.1 Paper A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
8.2 Paper B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
8.3 Paper C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
8.4 Paper D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

II Appended papers 65
Part I

Overview and Summary

1
Chapter 1

Introduction

A comfortable sound and vibration environment in passenger trains is an important


quality factor. On the other hand, obtaining low noise and vibration levels is often
expensive and can add weight to the train. Manufacturers work hard to improve the
comfort standards of trains, while at the same time try to keep the costs down. The
design of rail vehicles is driven by a number of functional requirements e.g load carry-
ing capacity, acoustic and thermal insulation, good aerodynamics, interior and exterior
aesthetics. The components and parts used in trains have been and still being designed,
produced and assembled separately, each fulfilling different functions.
An alternate approach to this conventional method is multifunctional design. The
idea of multifunctional design is to design/select a component so that it can have mul-
tiple functionalities which can reduce the number of total components. For example, a
multifunctional wall or a floor gives more space, higher comfort, lower weight, acous-
tic and thermal insulation while at the same time providing mechanical protection.
During the design of a multifunctional panel, solutions have to be obtained for several
design criteria such as static and dynamic stiffness, thermal insulation, acoustic insu-
lation, partition thickness, weight and production costs. In the final design, the vari-
ous functional requirements should be met while keeping the material and production
costs low and avoiding overly complex structures.
One of the main outcomes of a multifunctional panel is the weight reduction. Weight
reduction leads to reduced fuel consumption and reduced wear of the equipment.
However, an early study showed that it is difficult to comply with the noise and vi-
bration requirements in rail vehicles using a light weight structure [3]. It is therefore
important to take into account the various aspects of the functions and the constraints
while designing a multifunctional panel. According to Carlsson [3], the various noise
transmission paths into a passenger compartment for both air-borne and structure-
borne noise emissions are as shown in Figure 1.1.

3
Sathish Kumar

Figure 1.1: Noise and vibration transmission paths in a railway compartment [3].

1 Air-borne noise through floors, walls, windows, roofs and auxiliary equipments
like fans, motors, gears, HVAC units.


2 Structure-borne noise from bogie, diesel engines.

Air-borne sound often originates from external sources and propagates into the
vehicle interior through the floor, wall panels and also through the holes in the body-
work, door seals, etc. Whereas, structure-borne sound is the result of mechanical vi-
brations propagating through the vehicle structure and eventually causing localised
displacements in air. The noise from wheel-rail region is often the major source of both
air-borne and structure-borne sound. On vehicles with underfloor diesel engines, noise
from the engine can be significant and so is the noise from air conditioning systems.
Vibrational energy from these sources can be transmitted to the passenger compart-
ment. Thus, the floor structure of a passenger compartment is expected to possess
good sound insulation properties. For most passenger trains, floating floors are ap-
plied to obtain sufficient noise reduction from the sources under the floor. In addition,
lightweight and thin floor designs are desired for increased weight reduction.
The inner flooring of a passenger compartment made of sandwich panels with alu-
minium face sheets and honeycomb core material have certain advantages over floor
panels made of wood. For this reason, traditional floor panels made of wood is being
replaced by light weight sandwich structures. Some of the advantages of these con-

4
Overview and Summary

structions are low weight, good moisture properties, fire resistance and high stiffness-
to-weight ratio etc. As mentioned earlier, one main limitation of such light weight struc-
ture is the poor sound insulation property. In the transportation industry, automotive,
aircraft and rail vehicles are treated with damping materials to reduce structure-borne
sound and its effectiveness mainly depends upon parameters such as materials, loca-
tion and size of the damping treatment. The basic principle of damping is to convert vi-
bratory energy into heat. Traditional damping treatments using viscous damping lay-
ers are typically of two types, unconstrained and constrained layer damping [8]. Due to
the shear deformation occurring in the visco-elastic layer, constrained layer damping
treatments are known to yield significantly larger system damping compared to uncon-
strained layer damping, for the same mass of damping material used [19]. The damp-
ing treatments are effective in reducing structural vibration and it is often thought that
reduced vibration results in noise reduction. However, when the radiation efficiency
increases more than the vibration reduction, the total noise radiated from the struc-
ture increases [8, 17, 18]. This increased radiation due to the damping treatment does
not provide the desired effects on sound insulation. The lack of effective predictive
methods for assessing the sound insulation effects due to added damping on complex
industrial structures leads to excessive use of damping materials. Examples are found
in the railway industry where sometimes the damping material applied per carriage is
more than one ton.
As product development is carried out in a fast pace today, there is a strong need
for validated prediction tools to assist during early design stages. There is a need for
computationally efficient models for calculating the sound insulation properties of the
floors accounting all practical design aspects of typical train structures e.g. orthotropic-
ity of the corrugated plates, short circuiting due to boundary conditions and the effect
of stiffeners. In this thesis, prediction models for sound transmission through inner
floor panels used in floor structures are developed and validated. The effect of damp-
ing treatment on the sound insulation of inner floors is investigated. Also, a sound
transmission model for rib-stiffened double-leaf partitions is developed and extended
to account for complex design aspects such as orthotropicity, sound absorption in the
cavity, stud stiffness and placement. The aim of this thesis is to provide simple analyt-
ical models of sufficient accuracy and speed to predict the sound insulation of parti-
tions. The models should be useful for quick parameter studies at early design stages
and to be used together with optimisation techniques for better sound proofing in en-
gineering structures like airplane fuselage and railway floors. A brief introduction to
these structures is presented in the following chapter.

5
Chapter 2

Floor constructions in trains

2.1 Floating floors

Floating floors are used in vehicles to reduce noise and vibration levels. There are sev-
eral studies on the successful use of floating floors for structure-borne noise in the ship
and aircraft industry [33, 32, 26, 4]. For passenger trains, floating floors are applied to
obtain a sufficient noise reduction from sources under the floor. A typical floating floor
construction consists of an inner floor isolated from the car body with discrete elastic
inner layers (studs) resting on supporting beams (stiffeners) as shown in Figure 2.1.

Figure 2.1: Floating floor arrangement.

For thermal and acoustic insulation, the cavity between the inner floor and the car
body framework’s corrugated plate is filled with glass or mineral wool. Traditionally
the inner floor is made of plywood, but in recent years metal and composite sandwich
panels are increasingly used. Among other properties, sandwich structures provide
solutions for lightweight and thin floor designs. A properly designed floating floor
should: (i) Provide a vibration isolation for the inner floor for high frequencies. (ii)
Reduce the acoustical power radiated by the inner floor. (iii) Reduce transmission of
air-borne sound into the car.

7
Sathish Kumar

2.2 Sandwich constructions and the sandwich effect

Sandwich structures due to their low density in combination with high flexural stiff-
ness are extremely popular in aerospace and marine applications where weight is a
major issue e.g. in commercial planes, pleasure boats, space shuttles and satellites. In
ground transportation they are increasingly found in cars, buses and trains. A classical
sandwich structure consists of two stiff, strong, thin face sheets bonded to either side of
a relatively thick, weaker, light weight core material. The faces are usually made from
a high performance material such as steel, aluminium or fibre composite, whereas the
core is usually a structural solid foam, balsa wood or honeycomb (this can again be
made of aluminium, kraft paper, etc). The structural properties of the face sheets and
the core are less significant as individual panels, but when glued together to form a
sandwich, they produce a structure of high stiffness and high strength-to-weight ratio,
a property which is of great interest to the industry.
The good stiffness properties of a sandwich construction can be illustrated by the
following example. A structure made up of a homogeneous material with a given
Young’s modulus and strength having unit width and thickness 0 t0 will have a cer-
tain bending stiffness which is normalised as 1. Then the beam is cut into two halves
of thickness 0 t/20 and a core material of thickness 0 2t0 is bonded between these two
halves. The corresponding stiffness and strength is 0 120 and 0 60 times more than the
homogeneous beam respectively. This is called the sandwich effect. The core material is
assumed to have a surface density much lower than the face sheets and therefore any
addition in weight to the structure is considered negligible.

Weight Flexural Bending


Rigidity Strength

t 1 1 1

t/2
2t ~1 12 6

t/2
4t ~1 48 12

Figure 2.2: The sandwich effect [42].

8
Overview and Summary

2.3 Honeycomb panels

A honeycomb panel is a lightweight sandwich panel with a honeycomb core. Honey-


comb cores can be made of kraft paper, aluminium alloy, glass/phenolic and Nomex R
.

R
In the transportation industry, honeycomb panels with Nomex cores are standard in
aircraft floor structures, while aluminium cores are increasingly common in railway
car body structures as well as for certain building elements. For example, in the rail-
way car bodies, honeycomb panels are used in door panels, but they are also replacing
conventional wooden interior floors due to their resistance to moisture and also to al-
low for electrical floor heating systems. Honeycomb cores can be manufactured in a
variety of cell shapes, but the most commonly used shape is the hexagonal shape as
shown in Figure 2.3. In this thesis, three different honeycomb panels with aluminium
alloy cores are investigated. The panels are differentiated by their core thickness and
core structures as shown Figure 2.3 with dimensions and properties as shown in Table
2.1. The panels with 6.4 mm cell were selected to represent a standard design solution
for railway floor structures and the panel with 19.2 mm was chosen to study the effect
of a softer core.

Figure 2.3: Honeycomb panel layout with idealised core.

9
Sathish Kumar

Honeycomb t f /tc /t f Surface density


cell size (mm) (mm) of HCP (kg m-2)
Panel 1 6.4 1.5/18/1.5 10.4
Panel 2 6.4 1/10/1 6.8
Panel 3 19.2 1/10/1 6.2

Table 2.1: Properties of panels tested.

2.4 Damping treatments


Damping treatments are used to dissipate vibrational energy to control vibrations and
noise. Traditional damping treatments using viscous damping layers are typically of
two types, unconstrained and constrained layer damping [8]. Constrained layer damp-
ing (CLD) treatments have provided an effective means to impart damping to the
structure [2]. Due to the shear deformation occurring in the visco-elastic layer, CLD
treatments are known to yield significantly larger system damping compared to un-
constrained layer damping, for the same mass of damping material used [19]. Though
these damping materials are effective to attenuate vibration, their effect on the air-
borne sound transmission is often limited due to the increased radiation efficiency
[8, 17, 18]. In this thesis, the damping treatment applied to the inner floor panels con-
sists of a 1 mm thick visco-elastic layer with a constraining layer.
Assuming the core to be weak, the bending stiffness of a sandwich plate is mainly
influenced by the face separation distance. Panel 1 having a face separation distance of
18 mm has a high initial flexural stiffness. Therefore, a 1 mm thick steel plate (surface
density 7.8 kg m-2) was used as a constraining layer for panel 1 and 1 mm thick alu-
minium plate (surface density 2.7 kg m-2) was used for panel 2 and panel 3 which had
relatively low initial bending stiffness. The damping materials can be easily glued to
the honeycomb panel as shown in Figure 2.4.

Figure 2.4: Constrained damping layer attached to a honeycomb panel.

10
Chapter 3

Sound transmission through


honeycomb panels

3.1 Wave propagation in sandwich structures


Sandwich structures have been the subject of many studies, a large amount of litera-
ture has been devoted to the development of theories to study their static and dynamic
behaviours using analytical and numerical methods [20, 15, 9, 10, 31, 29, 25, 34]. Acous-
tic analysis of sandwich panels using the classical method was pioneered by Kurtze
and Watters [20] and later by Dym and Lang [9]. Kurtze suggested that using a sand-
wich panel can provide better sound insulation between partitions than a homogenous
panel of equivalent mass. Kurtze’s model for sound transmission is based on simple
sandwich theory where the effect of core shear is included. Dym [9] improved the accu-
racy of Kurtze’s model and later corrected his model [10] by taking into consideration
the acoustic excitation on the lower face sheet of a sandwich panel. Apart from the
classical methods discussed above, statistical energy analysis (SEA) [25] is often used.
Commercial softwares like VA One based on SEA and NOVA based on Transfer Ma-
trix Method (TMM) can predict the sound transmission loss (STL) of sandwich panels
when the panel properties are available. A comprehensive assessment of the above
mentioned sandwich models has been performed by Wang et al. [40].
A more general approach is given by Nilsson [31] based on the bending of sand-
wich beams. The sound insulation property of a sandwich panel is governed by the
properties of the constituent materials, the boundary conditions and the wave mo-
tions in the panel. Nilsson’s method is based on classical equations of motion for wave
propagation, as well as theories for excitation and radiation from plates. The general
wave equation is used to describe the displacement in the core and the laminates are
treated as thin Kirchhoff plates while the influence of boundaries are not discussed.

11
Sathish Kumar

The wavenumbers in a sandwich plate is an indicator of the response of the plate to an


external excitation. In particular, the waves determining the lateral bending of the plate
is of importance, since these have the greatest influence on excitation and radiation. In
the low-frequency limit, the propagation constants are equal to
r
4 ω2 m
kB = . (3.1)
D
Where ω is the angular frequency, m is the surface density of the panel, k B and
D are the bending wavenumbers and the apparent bending stiffness of the sandwich
panel respectively. For increasing frequencies, the lateral motion of the structure can no
longer be described by bending alone, the shear and rotation of the core will influence
the deflection of the plate and the apparent bending stiffness of the plate. For thin
walled faces it can be assumed that the displacement of the facesheet is determined by
the flexural and longitudinal waves and the displacement in the thick core is due to
bending, rotation, shear and as well as longitudinal deflection.

θ
pr
pi
E1, t 1
y
E2, G2 , t 2
x
E1, t 1
pt

Figure 3.1: Geometry and material parameters of a sandwich panel with an isotropic
core.

For a sandwich panel with symmetric face sheets and an isotropic core as shown in
Figure 3.1, the wavenumbers can be derived as a solution to the following expression
[31],

F (k ) = (U1 − U2 ) (V3 − V4 ) − (U3 + U4 ) (V1 + V2 ) . (3.2)

The functions U and V are given in Appendix- Part A. A detailed account about
deriving these functions can be found in the original Reference [31]. The solutions for
k are obtained when F = 0.
Due to the manufacturing of honeycomb cores, the cell wall thickness doubles in the
direction of cell orientation and the actual shape of the honeycomb cell can be irregular
making the core anisotropic as shown in Figure 3.2.
For simplicity, the honeycomb core can be treated as orthotropic and the whole

12
Overview and Summary

(a) Regular (b) Irregular

Figure 3.2: Honeycomb core shapes.

panel can be treated as an orthotropic sandwich panel, if the frequency concerned is


not too high. The complex wavenumbers k Bx and k Bz in the x- and z- directions (also
called L- and W- directions) as shown in Figure 3.3 can be solved by using an aver-
aged value of the loss factor measured between 100 Hz and 1000 Hz and by applying
different values for the core shear modulus G2x and G2z for the x − and z− directions
respectively. Along with the geometrical properties, the parameters in Table 3.1 are
used to calculate the wavenumbers. The shear moduli in Table 3.1 is supplied by the
panel manufacturer.

Panels G2x (Pa) G2z (Pa) Loss factor Loss factor Poisson’s
HCP HCP + CLD ratio of core
Panel 2 430 × 106 220 × 106 0.005 0.04 0.33
Panel 3 201 × 106 54 × 106 0.005 0.04 0.33

Table 3.1: Parameters used for wavenumber calculations.

3.2 Sound transmission through orthotropic plates


For an orthotropic panel as shown in Figure 3.3, the lateral displacement is governed
by [17],

∂4 W ∂4 W ∂4 W
Dx + 2D xz + D z − ω 2 mW = jωΦ, (3.3)
∂x4 ∂x2 ∂z2 ∂z4
where Dx and Dz are the bending stiffness in x − and z− directions respectively and

Dxz can be approximated by Dxz ≈ Dx Dz .

13
Sathish Kumar

Figure 3.3: Geometry of a sandwich panel with an anisotropic core.

Dx and Dz can be defined as the bending stiffness of a honeycomb structure having


the same dynamic properties as a simple homogeneous beam at certain frequencies.
According to Heckl [17], for a given angle of incidence θ, the transmission coefficient
of an infinite orthotropic plate is given by

 !2  −2
2
cos2 φ

ωm cos θ  sin φ 4 4 

τ (θ, φ) = 1 + j 1− + 2 k0 sin θ , (3.4)
2ρ0 c0 k2Bx k Bz
p p
where k Bx = 4 ω 2 m/Dx and k Bz = 4 ω 2 m/Dz are the wavenumbers in x − and
z−directions respectively. For small loss factors, the damping can be included by re-
placing Dx and Dz in Equation 3.4 with Dx (1 + jη ) and Dz (1 + jη ) respectively [17].
Assuming the same loss factor for waves in all the directions in the panel, Equation 3.4
can be re-arranged as
" #2 " !#2 −1
 ωm cos θ k40 sin4 θ ωm cos θ k40 sin4 θ 
τ (θ, φ) = 1+η + 1− , (3.5)
 2ρ0 c0 k4φ,eqv. 2ρ0 c0 k4φ,eqv. 

where k φ,eqv. is the equivalent wavenumber in direction φ given by

1 sin2 φ cos2 φ
= + 2 . (3.6)
k2φ,eqv. k2Bx k Bz

The diffuse field sound transmission coefficient over all angles of incidence θ and trans-
mission φ is calculated by [36]

14
Overview and Summary

´ 2π ´ θlim
0 0 τ (θ, φ) sin (θ ) cos (θ ) dθdφ
τ= ´ 2π ´ θlim , (3.7)
0 0 sin ( θ ) cos ( θ ) dθdφ
where θlim , is the limiting angle above which it is assumed that no sound is incident
on the partition and according to Mulholland et al. [30] θlim can be limited to 78◦ . The
sound transmission loss is then calculated by
 
1
STL = 10 log10 . (3.8)
τ

3.3 Radiation loss factor


Radiation loss factor is a parameter indicating to what extent the vibration of a struc-
ture is damped due to radiation to the surrounding fluid. According to Cremer et al.
[8], the radiation loss factor can be defined as

ρ0 c0 σ
ηrad = , (3.9)
ωm
where ηrad is the radiation loss factor and σ is the radiation efficiency which has
been investigated by Beranek [2] and Maidanik [27]. According to Leppington et al.
[24], a better estimation of the averaged radiation efficiency of a rectangular plate can
be given as
   
a+b µ+1 2µ
σ̄ ≈ ln + 2 µ > 1,
πµkab (µ2 − 1) µ−1 µ −1


σ̄ ≈ kaH ( x ) µ = 1, (3.10)

q
σ̄ ≈ 1/ 1 − µ2 µ < 1,
p
where a and b are the width and length of the plate with a < b, µ = k B /k = f c / f
with f c being the critical frequency of the plate, and H ( x ) is a function of the panel
aspect ratio ( a/b) and in our case it can be approximated to 0.41 [24].
For an isotropic panel, there is one critical frequency and the radiation loss factor
can be applied as in Reference 31. But for orthotropic panels, there exists different crit-
ical frequencies depending on the transmission angle φ. For anisotropic panels, there
exists no expressions for calculating the wavenumbers and critical frequency as a func-
tion of φ. Since the honeycomb panel is treated as being orthotropic , the equivalent
critical frequency in the ribbon (x) and transverse (z) directions can be calculated using

15
Sathish Kumar

the equivalent wavenumber from Equation 3.6 and the related radiation loss factor can
be obtained.
The loss factor is normally measured in 1/3rd octave bands because of the large
fluctuations in the measurements. However, Feng and Kumar [13] showed that by us-
ing a measured loss factor in 1/3rd octave band, the sound transmission loss will be
underestimated around the critical frequency because of the extremely large radiation
loss factor concentrated in a very small frequency band as indicated in Equation 3.10.
As a consequence, the radiation loss factor calculated in narrow band has to be added
to the measured loss factor for better sound transmission loss predictions around the
critical frequency [13]. The total loss factor can be expressed as sum of the measured
loss factor and the radiation loss factor.

ηtotal = η0 + 2ηrad . (3.11)

16
Chapter 4

Sound transmission through


double-leaf partitions with periodic
stiffeners

The transmission loss of regular double-walls is extensively covered in the literature


[27, 6, 7, 21, 39, 22, 41]. For simpler structures like single- or double-leaf partitions
made of homogenous panels, the wave approach by Cremer et al. [8] and Fahy [11]
gives a good estimate of the sound transmission loss. The advantage of this method is
that it can be applied to a wide frequency range and takes dominant physical phenom-
ena into account, like the wave coincidence and mass law effects by relatively simple
calculations. A limitation of this wave approach is that it cannot be applied to non-
homogenous panels and for partitions with stiffeners [35] and hence cannot be applied
to typical engineering structures found in rail vehicles, aircraft and ships. Statistical
energy models of double-leaf partitions are frequently used due to their simplicity and
computational efficiency [6, 38, 5]. Novel work by Finnveden and Barbagallo [14] al-
lows a physically correct description of the wave transmission at lower frequencies
accounting for double-wall resonance, but models of rib-stiffened structures need to
be developed further. In particular, detailed effects of parameter variations of stud
stiffness and placement, stiffener geometry and layout are not readily available with
computationally efficient models.
In this thesis, space harmonic analysis is used to investigate rib-stiffened double-
leaf partitions. The structural response and acoustic pressure of a periodic structure
due to a harmonic excitation has been expressed in terms of a series of space harmon-
ics by Mead and Pujara [28]. In their work, the panel was modelled as a beam resting
on elastic supports resisting both transverse and rotational displacements. This model
predicted the structural response due to an acoustic pressure. However, it failed to

17
Sathish Kumar

Figure 4.1: Double-leaf partition with stiffeners and absorbing materials in cavity.

take into account the effect of the structural response on the acoustic system. Lee and
Kim [21] adopted the method of Mead and Pujara to a 2-D single plate with stiffeners
and included the effect of structural response on the acoustic system. The stiffeners
were modelled as periodically distributed lumped masses mounted to the two pan-
els, together with a series of uniformly distributed translational and rotational springs
and a detailed approach was presented to calculate the sound transmission loss of the
structure. As an extension of Lee’s work [21], Wang et al. [39] studied double-walled
rib-stiffened panels and following Lee, Wang disregarded many complicating features
(e.g. sound absorbing material in the cavity). In this thesis, Wang’s model has been
extended to include the effects of different fluid properties in the centre cavity which
accounts for the sound propagation through an absorbing material. Also, the model is
modified so that orthotropic sandwich panels can be included as a part of the double-
walled partition. Naturally, the same model can be simplified and used to investigate
the sound transmission properties of a double-leaf partition with standard sandwich
panels.

4.1 Space harmonic analysis - Isotropic panels


A double-walled structure can be modelled with periodic boundary conditions as shown
in Figure 4.2a. For simplicity, the system is considered to be infinite in the direction
of the stiffeners and the incident of harmonic wave is limited only to the horizontal
plane. The panels are considered isotropic with thickness hi , and mass per unit area
m pi of panel i (where i = 1, 2). The two panels are linked with five vertical “C” shaped
aluminum channels. These stiffeners are modelled as periodically spaced translational
0
and rotational springs and their weights M , are equally distributed to the two pan-
els. A stiffener of depth l, web thickness to , Young’s modulus E can be modelled as a
spring of translational stiffness Kt0 = Et0 /l per unit length of stiffener [39]. Similarly,
the rotational stiffness is estimated by considering the web bending of the stud and is

18
Overview and Summary

given by Kr0 = EIz /l where Iz is the second moment of area about the z-axis. For “C”
shaped stiffeners studied in this thesis, Iz = t3o /12.

(a) Periodic representation (b) Notation used in the analysis

Figure 4.2: Cross-section of a double-leaf panel with stiffeners.

The plane wave pi , incident on the periodic panel with stiffeners induces a reflected
wave pr , a transmitted wave pt and the panel motion. Since the structure is periodic,
the system response is also expected to be periodic [28], the motion of each panels can
be expressed in terms of space harmonic expansions. The plane wave is assumed to
incident along the x-y plane at an angle θ as shown in Figure 4.2b and the incident
wavenumber components k x and k y are given as

k x = k sin θ , k y = k cos θ , (4.1)

where k = ω/c0 and ω is the angular frequency (2π f ) and c0 is the speed of sound in
air. Similarly, the velocity potentials in the incident, cavity and the transmitted space
can be expressed in terms of space harmonic series [28]. The wavenumber in the y-
direction in, k yn is given by the expression
s
 2
2 2nπ
k yn = (k) − k x + . (4.2)
L

Since the cavity is filled with an absorbing material, the wavenumber in the y-direction
in the cavity, k ync is given by the expression

19
Sathish Kumar

s
 2
2 2nπ
k ync = (k c ) − k x + . (4.3)
L

where k c is the complex wavenumber in the absorbing material and is related to the
dynamic density ρc (ω ) and the dynamic bulk modulus Kc (ω ) of the absorbing material
as

s
ρc (ω )
kc = ω . (4.4)
Kc ( ω )
The dynamic density and the dynamic bulk modulus are obtained through Allard-
Champoux model for rigid frame materials [1]

    
1 σ ρo f
ρ c ( ω ) = ρ0 1+ G1 , (4.5)
j2π ρ0 f σ

! −1
γ−1
Kc (ω ) = γP0 γ− . (4.6)
1 + 1/j8πNpr (ρo f /σ )−1 G2 (ρ0 f /σ )


p
The functions G1 (ρ0 f /σ) and G2 (ρ0 f /σ) are given by G1 (ρ0 f /σ) = [1 + jπ (ρ0 f /σ )]
 
and G2 (ρ0 f /σ ) = G1 (ρ0 f /σ) 4Npr , where ρ0 is the density of air, σ is the static air
flow resistivity, γ is the specific heat ratio, P0 is the air equilibrium pressure and Npr
is the Prandtl number. A real part of the wavenumber greater than unity indicates a
lower speed of sound through the absorbing material and a negative imaginary part
indicates the dissipation of energy. Applying the boundary conditions at the fluid-
panel interface, the modal amplitudes can be found in terms of the amplitudes of the
space harmonics. The amplitudes of the space harmonics can be found by applying the
principle of virtual work as explained in Appendix-Part B. Making use of the coupling
relations between the modal amplitudes of waves in air and the flexural motion of
panel, the following relations are obtained.

20
Overview and Summary

  
2ρ 2jk ync H
 4 jω jω 2 ρ
c 1 + e
 D1 k x + 2mπ − m p1 ω 2 + 0
−    α1,m
L k yn k ync 1 − e 2jk ync H

+∞ +∞
! "  #  
 0 0
 0 2nπ 2mπ
2
+ Kt − ω M ∑ α1,n + Kr ∑ α1,n k x + L kx +
L
n=−∞ n=−∞
+∞
! "
+∞   #  (4.7)
0 0 2nπ 2mπ
− Kt ∑ α2,n − Kr ∑ α2,n k x + L kx +
L
n=−∞ n=−∞

2jω 2 ρc e jkync H 2jωρ I, f or m = 0,
0
+   α2,m =
k 1 − e2jkync H
ync
0, m 6= 0

 
4 jk yn H jk ync H
jω 2 ρ 2jω 2 ρ

 D2 k x + 2mπ 0e ce
− m p2 ω 2 + −   − α2,m
L k yn
k ync 1 − e 2jk ync H

+∞ +∞
!  " #  
 0 0
 0 2nπ 2mπ
+ Kt − ω 2 M ∑ α2,n + Kr ∑ α2,n k x + L kx +
L
n=−∞ n=−∞
+∞
! "
+∞  #
   (4.8)
0 0 2nπ 2mπ
− Kt ∑ α1,n − Kr ∑ α1,n k x + L kx +
L
n=−∞ n=−∞
 
jω 2 ρc 1 + e2jkync H
+   α1,m = 0.
2jk ync H
k ync 1 − e

These equations are solved for αi,m which can be used to find the modal amplitudes.
The power transmission coefficient is then calculated by


∑nn=+ 2 Re k

=−∞ | ξ n | yn
τ (θ ) = . (4.9)
| I |2 k y

The diffuse field sound transmission coefficient over all angles of incidence θ within
the incident plane is calculated by [36]

´ θlim
0 τ (θ ) sin (θ ) cos (θ ) dθ
τ= ´ θlim . (4.10)
0 sin ( θ ) cos ( θ ) dθ

21
Sathish Kumar

4.2 Space harmonic analysis - Sandwich panels


The double-leaf partition with isotropic panels discussed in the previous section could
be ideal for walls in building constructions, but not as floors e.g. in trains because of
their poor load carrying capacity. By replacing the homogenous panels with sandwich
materials, the load carrying capacity of such partitions can be increased significantly
without adding additional weight. The model presented in the previous section has
been extended to replace the isotropic panels with sandwich panels. Specifically, sand-
wich panels with different properties in the orthotropic directions are studied in this
section.
Following the methodology for isotropic plates in the previous section, the double-
leaf partition as shown in Figure 4.3a with orthotropic sandwich panels connected by
stiffeners can be modelled with periodic boundary conditions . For simplicity, the sys-
tem is considered to be infinite in the direction of the stiffeners and the angle of inci-
dence of harmonic wave is limited only to the horizontal plane. The panels are consid-
ered orthotropic with thickness hi , mass per unit area m pi and Young’s moduli Exi , Ezi
of panel i (where i = 1, 2 and x and z are the orthotropic directions). The stiffeners are
modelled as periodically spaced translational and rotational springs with their masses
0
M , equally distributed on the two panels. The translational and rotational stiffness of
the beam shaped stiffeners can be estimated using the formulas in the previous section.

(a) Periodic representation (b) Notation used in the analysis

Figure 4.3: Cross-section of a double-leaf panel with stiffeners.

The plane wave pi , incident on the periodic panel with stiffeners induces a reflected
wave pr , a transmitted wave pt and the panel motion. Since the structure is periodic,

22
Overview and Summary

the system response is also expected to be periodic [28], the motion of each panels can
be expressed in terms of space harmonic expansions. The plane wave is assumed to
incident along the x-y plane at an angle θ as shown in Figure 4.3b and the incident
wavenumber components k x , k y and k z are given as

k x = k sin θ cos φ , k y = k cos θ , k z = k sin θ sin φ , (4.11)

where k = ω/c0 and ω is the angular frequency (2π f ) and c0 is the speed of sound in
air. The wavenumbers in the y-direction, k yn and k ync are given by the expressions

s
 2
2 2nπ
k yn = (k) − k x + − ( k z )2 and (4.12)
L

s
 2
2 2nπ
k ync = (k c ) − k x + − ( k z )2 . (4.13)
L

Applying the boundary conditions at the fluid-panel interface, the relationship be-
tween modal amplitudes and the amplitudes of the space harmonics are obtained. Ap-
plying the principle of virtual work, making use of the coupling relations between the
modal amplitudes of waves in the fluid and the flexural motion of panel, the following
relations are obtained.

"  4  2
2mπ 2mπ
Dx1 kx + + Dxz1 kx + (k z )2 + Dz1 (k z )4 − m p1 ω 2
L L
 
jω 2 ρc 1 + e2jkync H +∞
!
jω 2 ρ  0 0

∑ α1,n
0
+ −    α1,m + Kt − ω 2 M
k yn k ync 1 − e 2jk ync H n=−∞

+∞ +∞
"  #   !
0 2nπ 2mπ 0
+Kr ∑ α1,n k x + kx + − Kt ∑ α2,n (4.14)
n=−∞ L L n=−∞
+∞
"  #  
0 2nπ 2mπ
−Kr ∑ α2,n k x + kx +
n=−∞ L L

2jω 2 ρc e jkync H 2jωρ I, f or m = 0,
0
+   α2,m =
k 1 − e2jkync H
ync
0, m 6= 0

23
Sathish Kumar

"  4  2
2mπ 2mπ
Dx2 kx + + Dxz2 kx + (k z )2 + Dz2 (k z )4 − m p2 ω 2
L L

+∞
!
jω 2 ρ0 e jkyn H 2jω 2 ρc e jkync H  0 0

+
k yn
− 
2jk ync H
 − α2,m + Kt − ω 2 M ∑ α2,n
k ync 1 − e n=−∞

+∞ +∞
"  #   !
0 2nπ 2mπ 0
+Kr ∑ α2,n k x + kx + − Kt ∑ α1,n (4.15)
n=−∞ L L n=−∞
+∞
"  #  
0 2nπ 2mπ
−Kr ∑ α1,n k x + kx +
n=−∞ L L
 
jω 2 ρc 1 + e2jkync H
+   α1,m = 0.
2jk ync H
k ync 1 − e

Where, Dxi and Dzi are the bending stiffness of the sandwich panels in two orthotropic

directions and Dxzi can be approximated by Dxzi ≈ Dxi Dzi . For a honeycomb panel,
Dxi and Dzi can be obtained using the method described in Chapter 3.
These equations are solved for αi,m which can be used to find the modal amplitudes.
The power transmission coefficient is then calculated by


∑nn=+ 2

=−∞ | ξ n | Re k yn
τ (θ, φ) = . (4.16)
| I |2 k y

The diffuse field sound transmission coefficient over all angles of incidence θ and trans-
mission φ is calculated by [36]

´ 2π ´ θlim
0 0 τ (θ, φ) sin (θ ) cos (θ ) dθdφ
τ= ´ 2π ´ θlim . (4.17)
0 0 sin (θ ) cos (θ ) dθdφ

Where θlim , the angle above which it is assumed that no sound is received is limited to
78◦ [30]. The sound transmission coefficients from Equations 4.10 & 4.17 can be used
to calculate the sound transmission loss by
 
1
STL = 10 log10 . (4.18)
τ

The model in this thesis is developed for a double-leaf partition made of two or-
thotropic sandwich panels but naturally, isotropic material properties can also be ap-
plied.

24
Overview and Summary

4.3 Convergence
Since the solution to this space harmonic analysis is obtained in a series form, a conver-
gence check is performed to establish the number of terms required for accurate sound
transmission loss predictions.

4.3.1 Traditional convergence criterion


The sound transmission loss is calculated using a simple algorithm adding one term
at a time to the assumed expansion solution for selected frequencies (e.g. f = 200 Hz,
1000 Hz, 5000 Hz, & 10000 Hz). The solution is considered to have converged when the
STL difference between two successive calculations is less than a preset error of 0.01
dB. Figure 4.4 shows an example of how the calculated STL changes with increasing
the number of terms and the convergence associated with the selected frequencies. At
low frequencies few terms are enough to ensure convergence, but the number of terms
required increases with frequency. Following Lee and Kim [21], it can be concluded that
when a solution converges at a given frequency, it converges for all lower frequencies.

70
Convergence line

60

50
STL (dB)

40
200 Hz
1000 Hz
30 5000 Hz
10000 Hz
20

10

0
0 10 20 30 40 50 60 70 80 90 100
Number of Terms

Figure 4.4: Convergence check for different frequencies.

The standard procedure is to find the number of terms required at the highest fre-
quency of interest and thereafter calculate the sound transmission for all other frequen-
cies with the same number of terms. The cost of using a fixed number of terms for the
series expansion is the high computation time. To reduce computation time, a new cri-
terion is developed here for selecting the space harmonic terms for quick convergence.

25
Sathish Kumar

4.3.2 New convergence criterion

The basis of the new convergence criterion is that the pressure field at a given fre-
quency can be expressed with a limited number of propagating and evanescent waves.
At low frequencies this number of propagating waves is low, but it increases with fre-
quency and a larger number of terms is needed for convergence. In this new method
only the space harmonics corresponding to the propagating and a few evanescent
waves are included in the solution for each frequency. Neglecting cavity absorption,
propagating and evanescent waves can be separated by checking the wavenumbers
in the y-direction. For double-leaf partitions with isotropic panels, the wavenumber
in the y-direction is given by Equation 4.2. When ω/c0 >| k x + 2nπ/L | the wave is
propagating and when ω/c0 <| k x + 2nπ/L | the wave is evanescent (wavenumber
has only an imaginary part). This is illustrated in Figure 4.5 where the wavenumber
k yn is plotted against the space harmonics at selected frequencies. Since the evanescent
waves do not possess a real part, their amplitude is shown as 0.

200

180 1000 Hz
5000 Hz
160 10000 Hz
Space Harmonics
140 corresponding to
Propagating waves
120
real (k yn)

100 Space Harmonics


corresponding to
80 Evanescent waves

60

40

20

0
−40 −30 −20 −10 0 10 20 30 40
n

Figure 4.5: Real k yn plotted against the space harmonic terms, θ = 0◦ (Old criterion).

For a given frequency f , angle of incidence θ and stiffener separation distance L,


the propagating waves are limited to the region nmin ≤ n ≤ nmax where

Lf
nmin = − (1 + sinθ ) , (4.19)
c0

Lf
nmax = + (1 − sinθ ) . (4.20)
c0

26
Overview and Summary

By evaluating α1,n and α2,n in Equations 4.7 & 4.8 with n ranging from nmin to
nmax , includes only the propagating waves in the solution. Since the panels are closely
spaced, some evanescent waves do not decay sufficiently and have considerable am-
plitudes when reaching the second panel. The space harmonic terms corresponding
to these evanescent waves cannot be ignored and therefore, a number of evanescent
waves (∆n) should be included in the solution (n = nmin − ∆n to nmax + ∆n). A nu-
merical investigation revealed that evanescent waves with amplitude decays of more
than 60 dB when reaching the second panel, do not significantly affect the sound trans-
mission loss and therefore need not be included in the calculation. The old and the new
criterion for convergence can be illustrated in Figures 4.6 & 4.7.

200 200
Propagating waves Propagating waves

150 150
real (kyn)

real (kyn)

Evanescent waves Evanescent waves


100 100

50 50

0 0
10000 10000
8000 8000
6000 20 6000 20
4000 10 10
4000
0 0
2000 2000 −10
−10
100 −20 100 −20
Frequency (Hz) Frequency (Hz) Space Harmonic terms
Space Harmonic terms

(a) Old criterion (b) New criterion

Figure 4.6: Real k yn plotted against the space harmonics and frequency, θ = 0◦ .

100
100 Propagating waves
Propagating waves
80
80
60
real (kyn)

Evanescent waves
real (kyn)

60
40 Evanescent waves
40
20
20
0
90 0
80
70 80
60 20
50 60
40 20 10
30 40
10 0
20 0 20
10 −10 −10
0 −20 0 −20
Angle of Incidence (θ) Angle of Incidence (θ) Space Harmonic terms
Space Harmonic terms

(a) Old criterion (b) New criterion

Figure 4.7: Real k yn plotted against the space harmonics and angle of incidence, f =
5000 Hz.

In Figure 4.6, the wavenumbers are shown as a function of frequency with normal
incidence, wave heading (θ = 0◦ ), whereas in Figure 4.7, the wavenumbers are plotted
at 5000 Hz as a function of wave heading angle. For the new criterion, only the space
harmonics which are included in the solution are plotted against the real part of the

27
Sathish Kumar

wavenumber k yn for different frequencies. The non-zero wavenumbers in the plot cor-
respond to propagating waves and the others corresponds to evanescent waves. The
number of evanescent waves included in the solution, ∆n = 5.
And for orthotropic plates, from Equation 4.12, it can be determined that the waves
are propagating when ω/c0 >| (k x + 2nπ/L) + k z | and evanescent when ω/c0 <|
(k x + 2nπ/L) + k z |. Hence, for a given frequency f , angle of incidence θ, and angle of
propagation in the plate φ and stiffener separation distance L, the propagating waves
are limited to the region nmin ≤ n ≤ nmax where
q 
Lf 2
nmin =− 1 − (sinθsinφ) + sinθcosφ , (4.21)
c0
q 
Lf 2
nmax =+ 1 − (sinθsinφ) − sinθcosφ . (4.22)
c0

28
Chapter 5

Measurements

5.1 Loss factor

Loss factor is a good measure of the structural damping present in a system. From the
several methods available, a simple and robust decay method was used to measure
the loss factor of the test panels in 1/3rd octave bands. The test panel was suspended
in springs to achieve free-free boundary conditions minimizing boundary losses apart
from radiation. For the same reason, impulse excitation was used to avoid unwanted
external damping due to a shaker mounting. The decay method uses the fact that free
vibrations decay with time and by measuring the reverberation time of the structure,
the loss factor can be calculated by [8]

2.2
η= , (5.1)
fn T
where f n is the frequency and T is the reverberation time in seconds. Here, random
accelerometer positions and excitation points were used to get a spatial average of the
measured vibration response and to reduce the influence from individual modes.
After the damping layers were attached, the reverberation time decreased signifi-
cantly making decay measurements difficult. Therefore, power injection method was
used to measure the loss factor of the structures with damping treatment. This method
is not ideal for complex built-up structures, but works very well for freely hanging
simple structures as in our case. For a point excited system, the input power is related
to the spatial average of the vibration velocity as P = Sm0 ωηv2∆ /2 [8]. This equation
can be re-arranged to calculate the loss factor, as

2P
η= , (5.2)
Sm0 ωv2∆

29
Sathish Kumar

where P is the input power and S, m0 , v2∆ the area, surface density and the mean
square vibration velocity of the test structure respectively.
In practice, using harmonic excitation, the loss factor is determined according to
Feng [12], as

− Img( G f a )
η= , (5.3)
Sm0 Gaa
with G f a being the cross-spectrum of the excitation force and response acceleration
and Gaa being the power spectrum of acceleration.

5.2 Sound reduction index


The sound reduction index (SRI) is a measure of the sound insulation of a partition and
was measured using the sound intensity technique according to ISO 15186 − 1 : 2000.
The test partition was mounted between a reverberation room and an anechoic room
as shown in Figure 5.1. Four loudspeakers were used to generate white noise in the
reverberation room in the frequency range 100 Hz to 5000 Hz to create a diffuse sound
field. A condenser microphone attached to a rotating boom was used to measure the
average sound pressure level inside the reverberation room. The sound intensity was
measured on a plane parallel to the panel surface facing the anechoic room.

Anechoic Room
Anechoic Room Reverberation Room
Reverberation Room
Test Panel
Test Panel

Rotating
Boom

Loud-
speaker

Intensity probe
Intensity probe Rotating Boom Loud Speaker

PC &
Signal Generator Amplifier

Figure 5.1: Setup for sound reduction index measurement.

The panel was mounted in such a way that flanking transmission was minimised

30
Overview and Summary

thus ensuring that the sound transmission was only through the test panel. Then the
air-borne sound reduction index was calculated as given below

R I = L P − 6 − ( L I + 10 log (Sm /S)) . (5.4)

Where R I is the sound reduction index, L P is the sound pressure level in the rever-
beration room, L I is the sound intensity measured in the anechoic room, Sm is the mea-
sured surface area and S is the area of the test specimen. Weighted (apparent) sound
reduction index RW 0 , a single numbered quantity used to describe the sound insulation

of partitions was calculated according to ISO 717 − 1 : 2006.

5.3 Vibration velocity level


Vibrations in a structure can be quantified by measuring the average vibration veloc-
ity levels. The influence of damping treatments on the vibration of a structure can
be seen by comparing the vibration velocity levels of the structure before and after
the damping treatment. To measure the vibration velocity, the test panel was fixed
between the reverberation room and anechoic room as in the sound reduction index
setup. Eight piezoelectric accelerometers were placed on the panel surface facing the
anechoic room. Vibration velocity was measured on the panel surface for an acoustic
excitation from the reverberation room. The measurement was repeated for different
accelerometer positions. By taking a spatial average of the vibration velocity from all
accelerometer positions, the vibration velocity level was calculated using the expres-
sion below
!
<v>
Lv = 20log10 , (5.5)
vre f

where Lv is the vibration velocity level, < v > is the average panel normal velocity
and vre f = 10−9 m s-1 is the reference value for vibration velocity.

5.4 Bending stiffness


For homogeneous panels the bending stiffness is frequency independent, but for mul-
tilayer panels, like honeycomb panels, the bending stiffness becomes frequency depen-
dent. Simplified modal tests of freely hanging honeycomb beams were made to mea-
sure their natural frequencies and to calculate their apparent bending stiffness. Sections
(2050 mm x 70 mm) of honeycomb panel 2 and panel 3 were cut out in W-direction (Fig-

31
Sathish Kumar

ure 2.3) and suspended using rubber strings to achieve free-free boundary conditions.
The measurement setup is shown in Figure 5.2.

Charge Amplifier
Rubber Strings

PC & DAQ
Beam Accelerometer

Impact Hammer

Figure 5.2: Setup for bending stiffness measurement.

An impact hammer was used to excite the beam and an accelerometer was used
to measure the response. The inertial effects due to accelerometer mass was neglected
since the mass of the accelerometer was too small (2.4 grams) compared to the test
beams (975 grams and 889 grams). Different accelerometer positions were selected and
measured to make sure to find all relevant modes. From the measured natural frequen-
cies the bending stiffness was calculated by [42]

ωn m0 L4x ωn m0 L4z
Dx = , D z = , (5.6)
α4n α4n
where ωn is the natural angular frequency, L the length of the beam, m0 the beam
surface density, n is the order of the natural frequency and α1 = 4.73, α2 = 7.85, α3 =
11.00 , α4 = 14.14 , αn = nπ + π/2 for n > 4 . From the measured bending stiffnesses,
the flexural wavenumbers of the beams were calculated from the relations
s s
4 ωn2 m0 4 ωn2 m0
kx = , kz = . (5.7)
Dx Dz

32
Chapter 6

Results and discussion

6.1 Honeycomb panels

The loss factor of the honeycomb panels were measured using the decay method as
explained in Section 5.1. The measured loss factors for all three honeycomb panels
without and with the damping treatment are shown in Figure 6.1.

0.1 0.1
Panel 1 Panel 2
0.09 Panel 1 + CLD 0.09 Panel 2 + CLD

0.08 0.08

0.07 0.07

0.06 0.06
Loss Factor

Loss Factor

0.05 0.05

0.04 0.04

0.03 0.03

0.02 0.02

0.01 0.01

0 0
100 125 160 200 250 315 400 500 630 800 1000 100 125 160 200 250 315 400 500 630 800 1000
Frequency (Hz) Frequency (Hz)

(a) Panel 1 (b) Panel 2

0.1
Panel 3
0.09 Panel 3 + CLD

0.08

0.07

0.06
Loss Factor

0.05

0.04

0.03

0.02

0.01

0
100 125 160 200 250 315 400 500 630 800 1000
Frequency (Hz)

(c) Panel 3

Figure 6.1: Measured loss factor of the honeycomb panels without and with damping
treatment.

33
Sathish Kumar

From the figures it can be concluded that there is a significant increase in the struc-
tural loss factor with damping treatment for all the three panels. The decay measure-
ments were limited to a frequency of 1000 Hz above which, the reverberation time was
too short to get a useful result. Due to the previously mentioned problem of obtaining
loss factor results using the decay method with the damping treatment, power injection
method was used later. The principal use of damping treatments is to reduce vibrations
in the structure. To see the effect of damping treatments on vibration attenuation, the
vibration velocity level of the panels were compared without and with the damping
treatment. Figures 6.2a, 6.3a & 6.4a show the averaged vibration velocity level of the
panels to an acoustic excitation without and with the damping treatment. A diffuse
sound field was created in the reverberation room and it was ensured that the acoustic
excitation levels were same for all panels and measurements. As expected, the damp-
ing treatments reduce the vibration levels for all three panels. Figures 6.2b, 6.3b & 6.4b
shows the vibration level difference of the panels before and after the treatment.

120 14
Panel 1
Panel 1 + CLD
12
110
Vibration Velocity Level (dB)

10
100

8
∆ Lv (dB)

90
6

80
4

70
2

60 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Lv without and with damping (b) ∆Lv

Figure 6.2: Measured vibration velocity level (Lv ) of panel 1.

120 14
Panel 2
Panel 2 + CLD
12
110
Vibration Velocity Level (dB)

10
100

8
∆ Lv (dB)

90
6

80
4

70
2

60 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Lv without and with damping (b) ∆Lv

Figure 6.3: Measured vibration velocity level (Lv ) of panel 2.

34
Overview and Summary

120 14
Panel 3
Panel 3 + CLD
12
110
Vibration Velocity Level (dB)

10
100

∆ Lv (dB)
90
6

80
4

70
2

60 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Lv without and with damping (b) ∆Lv

Figure 6.4: Measured vibration velocity level (Lv ) of panel 3.

The sound reduction index of the panels was measured using the sound intensity
technique as explained in Section 5.2. The procedure was repeated after the damping
treatments were applied. The measured sound reduction index of panel 1 without and
with the damping treatment is shown in Figure 6.5. Just as observed in the loss factor
results, the damping treatment increased the sound reduction index of the panel. How-
ever, this increase is not entirely due to the damping treatment. A part of this increase
is due to the mass added to the panel due to the damping treatment (masslaw).

45
Panel 1
40 Panel 1 + CLD

35
Sound Reduction Index (dB)

30

25

20

15

10

0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.5: Measured sound reduction index of panel 1 without and with damping.

For this reason, mass normalisation was done to remove any effect of added mass
on the results. This was done by subtracting the mass contribution ∆ = 20 log(m0 /m00 )
from the measured sound reduction indices in each frequency bands, where m0 and m00
are the mass densities with and without damping layers attached. The measured sound

35
Sathish Kumar

reduction index of panel 1 compared to the mass normalised sound reduction index
with damping treatment is shown in Figure 6.6a. The result shows that the damping
treatment has very little influence on the sound reduction index of the panel. Similarly,
the sound reduction index of panel 2 and panel 3 are mass normalised and the results
are shown in Figure 6.6b and Figure 6.6c respectively. As observed for panel 1, the
damping treatment on panel 2 has very little influence. On the other hand, for panel 3
a significant increase in sound transmission loss can be seen above 500 Hz.

45 45
Panel 1 Panel 2
40 Panel 1 + CLD 40 Panel 2+ CLD

35 35
Sound Reduction Index (dB)

Sound Reduction Index (dB)


30 30

25 25

20 20

15 15

10 10

5 5

0 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Panel 1 (b) Panel 2

45
Panel 3
40 Panel 3 + CLD

35
Sound Reduction Index (dB)

30

25

20

15

10

0
125 250 500 1000 2000 4000
Frequency (Hz)

(c) Panel 3

Figure 6.6: Measured sound reduction index of the honeycomb panels without and
with damping treatment (mass normalised).

To investigate and quantify this influence, weighted sound reduction index RW0 ,a

single number quantity characterising the sound insulation of a partition was calcu-
lated according to ISO 717 − 1 : 2006 and is presented in Table 6.1. For panel 1, the
damping treatment increases the sound insulation by 5 dB whereas the influence from
added mass is 4.9 dB which shows that the increased losses does not increase sound
insulation. On the other hand for panel 3, the sound insulation increases by 8 dB out
of which only 3.9 dB is due to the added mass, clearly showing that damping treat-
ment increases sound insulation. From Table 6.1, it can also be inferred that without
the damping treatment, panel 2 has a weighted sound reduction index 3 dB more than
panel 3. But, once the damping material is attached, both panels have the same sound

36
Overview and Summary

insulation rating. It should be noted that the only structural difference between panel
2 and 3 is the honeycomb cell size affecting the shear stiffness of the core.

Panels HCP HCP + CLD ∆(dB)


R0w (dB) R0w (dB)
Panel 1 25 30 4.9
Panel 2 22 27 3.6
Panel 3 19 27 3.9

Table 6.1: Weighted sound reduction index of the honeycomb panels without and with
damping treatment.

The lowest frequency at which the wavelength of bending waves in plate is equal to
the wavelength in air is known as the critical frequency. For a single-leaf isotropic panel,
there exists one critical frequency f c whereas, for a honeycomb panel there exists a
range of critical frequencies between f cL and f cW (where the subscripts L and W are
the two principal orthotropic directions). This critical frequencies in the honeycomb
panels can be better illustrated using dispersion curves, where the wavenumbers of
the panels are plotted together with the wavenumbers in air.

Using the sandwich theory, the wavenumbers of the honeycomb panels were cal-
culated in the two orthotropic directions (L and W) and plotted in Figures 6.7a and
6.7b. In these figures, critical frequency is observed where the plate wavenumbers
(k L and kW ) intersect the wavenumbers in air (k air ). In Figure 6.7, the parallel lines are
the asymptotes to the wavenumbers. The lower asymptote (k lower ) corresponds to the
wavenumber for pure bending of the entire construction whereas, the upper asymp-
tote (k upper ) corresponds to the wavenumber of flexural waves propagating in one of
the laminate. It can be seen in Figure 6.7b that the wavenumbers in W-direction (kW )
matches with the wavenumbers in air (k air ) over a large frequency range (Figure 6.8).
This implies that the wavelength of the bending wave in the plate is close to the wave-
length of sound in air which leads to coincidence over the entire frequency range.

37
Sathish Kumar

2 2
10 10
Wave Number (1/m)

Wave Number (1/m)


1 1
10 kair 10 kair
klower klower
kupper kupper
kL kL
kw kw

125 250 500 1000 2000 4000 8000 125 250 500 1000 2000 4000 8000
Frequency (Hz) Frequency (Hz)

(a) Panel 2 (b) Panel 3

Figure 6.7: Wavenumbers in the panels (predicted) and air.

2
10
Wave Number (1/m)

1
10 kair
klower
kupper
kL
kw

2000 4000 6000


Frequency (Hz)

Figure 6.8: Extended critical frequency range, Panel 3.

Later, the panels were cut in the W-direction into beams (2050 mm x 70 mm) and
their bending wavenumbers were measured. The predicted wavenumbers are com-
pared with the measurements in Figure 6.9 and a good agreement is achieved. The
exact values of the measured and predicted critical frequencies for the two panels are
shown in Table. 6.2. The measured and predicted critical frequencies in W-direction
were found to be within a range of 2%.

38
Overview and Summary

Panels f cL Hz f cW Hz f cW Hz
prediction prediction measurement
Panel 2 792 888 904
Panel 3 851 3632 3556

Table 6.2: Critical frequencies of panel 2 and panel 3 in L- and W-directions.

2
10
Wave Number (1/m)

1
10

Air
0
Panel 2−Measured
10 Panel 3−Measured
Panel 2−Predicted
Panel 3−Predicted

125 250 500 1000 2000 4000


Frequency (Hz)

Figure 6.9: Measured and predicted bending wavenumbers.

Further, the vibration velocity level difference of the panel is compared with the
sound reduction index difference before and after the application of damping. A typi-
cal effect of damping treatments can be seen in Figure 6.10a. For panel 1, even though
significant vibration reduction is achieved with damping treatment, the sound radiated
from the structure is not reduced to the same extent. For panel 2, the increased sound
reduction index is approximately equal to the vibration reduction as seen in Figure
6.10b. On the other hand for panel 3 in Figure 6.10c, an interesting phenomenon is ob-
served. For panel 3, the increase in sound reduction index is more than the vibration
reduction at certain frequencies. The reason for this effect is the extended coincidence
range for panel 3 as observed before. It is well know that damping is more effective in
and around the critical frequency. Naturally, when damping treatments are applied to
a panel with a wide critical frequency range, their effect is greater than for a standard
panel.

39
Sathish Kumar

14 14
∆ Lv ∆ Lv
12 ∆ SRI 12 ∆ SRI

10 10

8 8
∆ (dB)

∆ (dB)
6 6

4 4

2 2

0 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Panel 1 (b) Panel 2

14
∆ Lv
12 ∆ SRI

10

8
∆ (dB)

0
125 250 500 1000 2000 4000
Frequency (Hz)

(c) Panel 3

Figure 6.10: Comparison of the difference in vibration velocity level and sound reduc-
tion index without and with the damping treatment.

The sound reduction index of the honeycomb panel was predicted using the model
described in Section 3. Figure 6.11 shows the measured SRI of panel 2 compared with
the predictions made using the measured structural loss factor averaged over fre-
quency. It can be seen that for panel 2, the predictions with the average loss factor holds
good for frequencies below critical frequency ( f < f c ) but at ( f = f c ) and above the
critical frequency ( f > f c ), the sound reduction index is under predicted. According to
Heckl [17], the radiation loss factor can have a significant contribution for structures
such as composite structures that are very lightly damped similar to those investigated
in this thesis. As a consequence, the radiation loss factor calculated in narrow band
frequency as in Equation 3.10 is added to the measured loss factor for better sound re-
duction index predictions around the critical frequency as shown by Feng and Kumar
[13].

40
Overview and Summary

45
Panel 2 − Measured
40 Panel 2 − Prediction

35
Sound Reduction Index (dB)

30

25

20

15

10

0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.11: Measured and predicted sound reduction index of panel 2 .

Figure 6.12a shows the predicted sound reduction index calculated with an added
radiation loss factor. A good agreement can be seen between the measured curve and
the predictions when the theoretical radiation loss factor is added. The same compari-
son for panel 3 is shown in Figure 6.12b.

45 45
Panel 2 − Measured Panel 3 − Measured
40 Panel 2 − with Radiation Lossfactor 40 Panel 3 − with Radiation Lossfactor

35 35
Sound Reduction Index (dB)

Sound Reduction Index (dB)

30 30

25 25

20 20

15 15

10 10

5 5

0 0
125 250 500 1000 2000 4000 125 250 500 1000 2000 4000
Frequency (Hz) Frequency (Hz)

(a) Panel 2 (b) Panel 3

Figure 6.12: Measured and predicted sound reduction index with added radiation
losses.

41
Sathish Kumar

6.2 Double-leaf partitions

The periodic model derived in Section 4.1 is used to examine the vibro-acoustic perfor-
mance of a double-leaf partition with homogenous aluminium panels and aluminium
stiffeners.The predicted transmission loss is then compared to the experimental data
from Legault and Atalla [23]. The parameters listed in Table 6.3 are used to describe the
partition and the sound absorbing material in the cavity. Assuming the air in the cavity
to be at standard temperature and pressure we have ρ0 = 1.21 kg m-3, Npr = 0.702,
and P0 = 101325 Pa and the wavenumbers in the cavity are obtained using Equation
4.4.

h1 (m) 1 × 10−3 η1 0.01 l(m) 50.8 × 10−3


h2 (m) 2 × 10−3 η2 0.01 to (m) 3.175 × 10−3
ρ1 (kg m-3) 2742 ν1 0.33 L(m) 508 × 10−3
ρ2 (kg m-3) 2742 ν2 0.33 H(m) 50.8 × 10−3
E1 (Pa) 70 × 109 co (m s-1) 342 Kt0 (Pa) 4.37 × 109
E2 (Pa) 70 × 109 ρ0 (kg 1.21 Kr0 (Pa) 3.67 × 103
m-3)
M(kg) 0.885 Porosity 91% Flow 17700
resistivity
(N s m-4)

Table 6.3: Dimensions and material data of panels and fibrous material [23].

Using the traditional convergence criterion, the number of terms required at the
highest frequency of interest is determined and the sound transmission is calculated
for all other frequencies with the same number of terms. From the convergence curve in
Figure 6.13a, it is found that at least 65 terms are needed for convergence at 10 kHz. The
sound transmission loss of the partition predicted using the traditional convergence
criterion is compared to the measured data in Figure 6.13b. The calculations are made
in narrow bands with a frequency step of 10 Hz and later averaged over 1/3rd octave
bands. From the figure it can be concluded that an overall agreement is achieved.

42
Overview and Summary

60
70
Convergence line

60 50

Sound Reduction Index (dB)


50
40
STL (dB)

40
200 Hz 30
1000 Hz
30 5000 Hz
10000 Hz 20
20

10 Measured
10
Predicted
n= −32 to +32
0 0
0 10 20 30 40 50 60 70 80 90 100 125 250 500 1000 2000 4000 8000
Number of Terms Frequency (Hz)

(a) Convergence analysis (b) Measured and predicted (traditional conver-


gence).

Figure 6.13: Convergence analysis and sound reduction index of the partition.

To reduce computation time, a new frequency dependent convergence scheme is in-


troduced as explained in Section 4.3.2. In this new scheme, the space harmonics corre-
sponding to all propagating waves and some evanescent waves which are considered
to be influential for sound transmission are included in the calculation. The propagat-
ing waves are limited within a region given by Equations 4.19 & 4.20. However, the
number of evanescent terms (∆n) included in the solution has to be estimated. Fig-
ure 6.14 shows the calculated sound transmission loss of the partition using the new
convergence criterion with different number of evanescent terms (∆n).

60

50
Sound Reduction Index (dB)

40

30

20

∆n = 2
10
∆n = 5
∆n = 10
0
125 250 500 1000 2000 4000 8000
Frequency (Hz)

Figure 6.14: Predicted sound reduction index using the new convergence scheme with
different ∆n.

It can be observed that the sound transmission results obtained with ∆n = 5 is very
similar to that obtained with ∆n = 10 but with a smaller computational expense. A

43
Sathish Kumar

numerical investigation revealed that when ∆n > 5, the evanescent waves decayed by
60 dB when reaching the second panel. Therefore, for the partition investigated here it
can be concluded that ∆n = 5 is enough to achieve numerical convergence. Figure 6.15
shows the measured sound transmission loss compared to the predictions obtained by
the traditional fixed term method and the new adaptive method. For this partition,
the new method takes one tenth of the computation time of the traditional technique
without compromising the accuracy of the result.

60

50
Sound Reduction Index (dB)

40

30

20

10 Measured
Traditional Convergence (n = −32 to +32)
New Convergence with ∆n = 5
0
125 250 500 1000 2000 4000 8000
Frequency (Hz)

Figure 6.15: Measured and predicted sound reduction index with different conver-
gence schemes.

Double-leaf partitions were constructed in the lab using aluminium honeycomb


sandwich panels and 2 mm thick aluminum panel. The two panels were linked with
five vertical “C” shaped aluminum channels, each spaced by a distance L of 505 mm
and the cavity was filled with mineral wool as shown in Figure 6.16b. The thickness of
the stiffeners was 3 mm, while their web length H measured 50 mm and their flange
length measured 30 mm. The stiffeners and the aluminum panels have a Young’s mod-
ulus of 70 × 109 Pa, Poisson’s ratio of 0.33, and a density of 2742 kg m-3. The honey-
comb panels discussed in the previous section (panel 2 and panel 3) were used to build
two different double-leaf partitions. The stiffeners were connected to the honeycomb
panels using glue as shown in Figure 6.16a and the aluminium panels were connected
to the stiffeners using evenly spaced screws (100 mm) mounted at the middle of the
flange. The properties of the aluminium panels, stiffeners and the mineral wool com-
mon for both the partitions is given in Table. 6.4 and the properties of the honeycomb
panels are as given in Table. 2.1 and Table. 3.1.

44
Overview and Summary

(a) Stiffener glued to the honeycomb (b) Partition with mineral wool but without aluminium panel
panel

Figure 6.16: Assembly of double-leaf partition with honeycomb panel.

h2 (m) 2 × 10−3 co (m s-1) 342 L(m) 505 × 10−3


ρ2 (kg m-3) 2742 ρ0 (kg m-3) 1.21 H(m) 50 × 10−3
E2 (Pa) 70 × 109 l(m) 44 × 10−3 Kt0 (Pa) 4.77 × 109
η2 0.01 to (m) 3 × 10−3 Kr0 (Pa) 3.57 × 103
ν2 0.33 Porosity 94% M(kg) 0.89
Flow 35000
resistivity
(N s m-4)

Table 6.4: Dimensions and material data of panels and mineral wool.

The sound transmission loss of the partitions were measured using the sound in-
tensity technique between 100 Hz and 5000 Hz as described in Section 5.2. Figure 6.17
shows the measured sound transmission loss of the two partitions. A difference in the
sound transmission loss can be seen in the higher frequencies. The lower transmis-
sion loss of the partition with honeycomb panel 3 is due to its softer core. The sound
transmission loss of the two partitions were calculated using the periodic model for
sandwich panels as described in Section 4.2 and the results are shown in Figures 6.18
& 6.19. The bending stiffness of the honeycomb panels used in the calculation were
obtained analytically using the wave propagation theory as described in Reference 31.

45
Sathish Kumar

60

50
Sound Reduction Index (dB)

40

30

20

10
Partition with HC panel 2 (6.4 mm core)
Partition with HC panel 3 (19.2 mm core)
0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.17: Measured sound reduction index of the two partitions.

Using the traditional convergence criterion, it was determined that at least 37 space
harmonic terms were necessary for the solution to converge at 5000 Hz . Whereas, us-
ing the new convergence criterion with ∆n = 5 , the computation time for the partitions
was reduced by a factor of 10. Figures 6.18 & 6.19 shows that the predictions using the
periodic model agree well with the measurements.

60

50
Sound Reduction Index (dB)

40

30

20

10
Measured
Prediction
0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.18: Measured and predicted sound reduction index of the partition with hon-
eycomb panel 2.

46
Overview and Summary

60

Sound Reduction Index (dB) 50

40

30

20

10
Measured
Prediction
0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.19: Measured and predicted sound reduction index of the partition with hon-
eycomb panel 3.

Modelling the stud stiffness


In this section, estimation of stiffness for beam-like studs are discussed. Such studs
are common in both aerospace and rail vehicle structures. The stiffness of the stud
is a critical parameter in sound transmission predictions. For a “C” shaped stud, it is
assumed that the flange, and particularly the web, possess a bending resilient character
which deforms under lateral force and hence acting as a spring between the two panels.
For the partitions described in the previous section, the translational stiffness per unit
length of the stud with thickness 3 mm is Kt0 = 4.37 × 109 Pa when calculated using
Kt0 = Et0 /l . If the stiffener is much thinner for example, with thickness 0.3 mm, this
expression leads to Kt0 = 0.43 × 109 Pa which by the order of magnitude is much stiffer
than reality. The reason is that the stiffness expression (Kt0 = Et0 /l ) assumes that
the deformation is mainly due to axial deformation in the stud web. This assumption
holds true when the stud thickness is high (in our case 3 mm), but when the stud
thickness is small, stud deformation mainly results from web bending and the above
expression grossly over estimates the stiffness. The translational stiffness of a flexible
connector accounting for web bending has been discussed by Legault and Atalla [23]
and is calculated by Ktb 0 = Et3 /3l L2 where E is the Young’s modulus, t is the web
0 F o
thickness, l is the depth, and L F is the length of the flange of the stiffener. For a stiffener
of thickness 0.3 mm, this expression leads to a stud stiffness of Ktb 0 = 2.27 × 104 Pa.

47
Sathish Kumar

(a) t0 = 3 mm (b) t0 = 0.3 mm

Figure 6.20: Influence of thickness on the translational stiffness of an aluminium stiff-


ener.

A double-leaf partition consisting of two identical 13 mm thick gypsum boards


separated by 65 mm steel studs of thickness 0.5 mm is investigated. The cavity be-
tween the boards is filled with mineral fibres. This partition corresponds to wall #12
in Reference 16. The parameters listed in Table 6.5 is used to describe the wall and the
mineral fibre. For a stiffener with thickness 0.5 mm, the translational stiffness becomes
0 = 2.15 × 105 Pa.
Ktb

70

60
Sound Reduction Index (dB)

50

40

30

20

10
Measured
Periodic model
0
125 250 500 1000 2000 4000
Frequency (Hz)

Figure 6.21: Comparison of the measured and predicted sound reduction index.

Figure 6.21 shows a comparison of the predictions, using the new convergence cri-
terion, to the test data from Reference 16 and it can be concluded that an overall agree-
ment is achieved. The space harmonics corresponding to all propagating waves and
a few evanescent (∆n = 3) waves were included in the calculation. For intermediate
stud thicknesses, the stiffness is rather complex but can be obtained using experiments
or through detailed stiffener vibration models [37].

48
Overview and Summary

h1 (m) 13 × 10−3 η1 0.02 l(m) 65 × 10−3


h2 (m) 13 × 10−3 η2 0.02 to (m) 0.5 × 10−3
ρ1 (kg m-3) 630 ν1 0.28 L(m) 406 × 10−3
ρ2 (kg m-3) 630 ν2 0.28 H(m) 65 × 10−3
E1 (Pa) 2 × 109 co (m s-1) 342 Kt0 (Pa) 2.15 × 105
E2 (Pa) 2 × 109 ρ0 (kg m-3) 1.21 Kr0 (Pa) 33.65
M(kg) 0.5 Porosity 93% Flow 11400
resistivity
(N s m-4)

Table 6.5: Dimensions and material data of panels and mineral fibre [16].

49
Chapter 7

Conclusions and future work

The measurements carried out during this thesis illustrates that constrained layer damp-
ing treatments increase the loss factors for all the honeycomb panels tested. As ex-
pected, the vibration velocity levels also decreased after damping treatment. The effect
on the sound transmission properties is more complex. For panels 1 and panel 2 the
damping treatment does not affect the sound transmission loss apart from the effects
of the added mass. However, for panel 3, the same damping treatment increases the
sound insulation due to the panel’s extended coincidence range as demonstrated by
the dispersion curves. It is known that damping has its most influence around the crit-
ical frequency where the transmission is resonant. Naturally, panel 3 with its extended
coincidence range had its sound insulation improved when the damping treatment
was applied. The weighted sound reduction index, R0w of panel 2 was 3 dB higher
than that of panel 3 but with the damping treatment applied both panels have the
same weighted sound reduction index (R0w ). In-short, the visco-elastic damping treat-
ment increases the sound insulation of an acoustically bad panel (in our case panel 3).
Whereas, the same treatment on a normal panel (panel 1 and 2) does not improve the
sound insulation apart from the effects of added mass.
Radiation losses can be a significant contributor to the total losses for lightly damped
composite structures. Since the measured loss factor is averaged over 1/3rd octave
bands, the abrupt increase of loss factor around the critical frequency is not captured in
the measurements. Therefore, the measured structural loss factor alone is not enough to
characterise the total losses of a light weight panel around the critical frequency. Apart
from the measured structural loss factor, the radiation loss factor calculated in narrow
bands must also be included in the model to accurately predict the sound transmis-
sion loss of lightweight panels like those discussed in this thesis. The sound reduction
index calculated using the wave propagation approach, combined with orthotropic
theory, agrees very well with the measurements.

51
Sathish Kumar

An analytical model is developed for double-leaf partitions with periodic stiffeners


and sound absorption material in the cavity. The stiffeners are modelled as uniformly
distributed lumped masses attached to the two panels with a set of translational and
rotational springs. The sound absorption in the cavity is characterised by an equiv-
alent fluid model. The vibro-acoustic response is derived by using a series of space
harmonics and applying the principle of virtual work. The sound transmission losses
of isotropic double-leaf partitions are calculated using the periodic model and the re-
sults are compared with published test data. The periodic model is later extended to
include orthotropic sandwich panels. The frequency dependent bending stiffness of
the orthotropic panels are obtained using sandwich theory. Two different double-leaf
partitions are built using an aluminium panel and two different honeycomb panels
and their sound transmission losses are measured and compared to predictions from
the periodic model. The results show that the periodic model using the space harmonic
expansion method provides a good estimate of the sound transmission loss for all the
systems analysed.
Since the solution to this method is obtained in a series form, computational accu-
racy needs to be balanced with computation cost as calculation times increases with
the number of terms. For improved computational efficiency, an adaptive criterion is
proposed to select the appropriate space harmonic terms which are considered to be
influential for sound transmission through the partition. This criterion results in quick
convergence of the solution. For the partitions studied in this thesis, the computation
time was reduced by a factor of at least 10. This decrease in computation time enables
to use this model for quick parametric studies in early design stages.
The stud stiffness plays an important role in the sound transmission of double-leaf
partitions. A partition was chosen to illustrate the importance of selecting the right ex-
pression for calculating the stud stiffness based on stiffener thickness. Two such models
for calculating the equivalent stud stiffness have been discussed. However, more work
is needed on developing simplified models to calculate the stud stiffness at higher fre-
quencies and studs with complex geometry. It is common to use visco-elastic mounts
in a floating floor assembly. The effects of visco-elastic mounts on the sound trans-
mission should be investigated. The periodic model developed in this thesis could be
used in conjunction with structural optimisation techniques for better sound proofing
in engineering structures like airplane fuselages and railway floors in the future.

52
Chapter 8

Summary of appended papers

8.1 Paper A
On Application of radiation loss factor in the prediction of sound transmission loss
of a honeycomb panel
Leping Feng and Sathish Kumar.
This article shows the importance of radiation loss factor in predicting the sound
transmission property of lightly damped panels, especially around the critical fre-
quency region. When the sound transmission properties of a panel is investigated,
the radiation loss factor is often neglected for panels unless the structural damping
is very less or when the sound transmission around the critical frequency of the panel
is of interest. This article shows that by including the theoretical radiation loss factor
in narrow bands to the measured structural loss factor, the sound reduction index of
honeycomb sandwich panels can be predicted with good accuracy around the criti-
cal frequency. The wave propagation constants are calculated as given by Nilsson [31]
and the sound reduction index of the panels are predicted using the orthotropic panel
theory. The predictions are validated through laboratory measurements.

8.2 Paper B
Predicting the sound transmission loss of honeycomb panels using the wave propa-
gation approach
Sathish Kumar, Leping Feng and Ulf Orrenius.
The wave propagation approach can be used to predict the sound transmission
properties of sandwich panels having an isotropic core. In this article, the wave propa-
gation approach is used to predict the sound transmission loss of honeycomb panels. A
honeycomb panel is a lightweight sandwich panel with a honeycomb core of hexagon

53
Sathish Kumar

cell. Due to the manufacturing process of honeycomb cores, the cell wall thickness
doubles in the direction of cell orientation and the actual shape of the honeycomb cell
can be irregular making the core anisotropic. As an approximation, the honeycomb
core can be treated as orthotropic and the whole panel can be treated as a orthotropic
sandwich panel. The wavenumbers are calculated for the two principal orthotropic di-
rections. The orthotropic panel theory is used to predict the sound transmission loss
of panels with two different core structures. Visco-elastic damping with constraining
layers (CLD) are attached to these panel and the sound transmission loss is calculated.
The predicted wavenumbers and the sound transmission loss of the panels agree well
with laboratory measurements.

8.3 Paper C
Modelling the sound transmission through rib-stiffened double-leaf partitions with
cavity absorption
Sathish Kumar, Leping Feng and Ulf Orrenius.
The sound transmission loss of infinite, periodically rib-stiffened double-leaf parti-
tions with different fluids in the cavity is investigated. The sound transmission model
for double-leaf partitions with periodic stiffeners has been extended to include sound
absorbing materials in the cavity. The vibro-acoustic response of the panels is expressed
as a series of space harmonics and virtual work theory has been successfully applied
for calculating the sound transmission loss of the partition. The sound propagation
in the cavity is defined by assuming an equivalent fluid model. The number of space
harmonic terms required for convergence of the sound transmission loss is studied. To
reduce the computation time, a new criterion is introduced to select the appropriate
space harmonics based on frequency and angle of incidence. This new criterion en-
sures quick convergence and reduced computation time. For the partitions analysed,
the computation time was reduced by a factor of ten. The model is verified through
comparison with experimental data available in the literature.

8.4 Paper D
Modelling the sound transmission through rib-stiffened sandwich double-leaf par-
titions using space harmonic analysis
Sathish Kumar, Leping Feng and Ulf Orrenius.
An analytical model is developed for the sound insulation of infinite, periodically
rib-stiffened double-leaf partitions consisting of sandwich panels. The two panels are

54
Overview and Summary

separated by “C” shaped stiffeners and the cavity between the panels is filled with
mineral wool. The sandwich panel investigated in this paper has a honeycomb core
which makes the panels orthotropic. The apparent bending stiffness of the sandwich
panels in the two orthotropic directions are derived using sandwich theory. The cavity
between the sandwich panels is filled with absorbing material and the sound propaga-
tion in the cavity is defined by assuming an equivalent fluid model. The vibro-acoustic
response of the panels is expressed as a series of space harmonics and virtual work
theory has been applied for calculating the sound transmission loss of the partition. A
new criterion is employed for quick convergence which reduces the computation time.
For the partitions analysed, the computation time was reduced by a factor of ten. The
predicted transmission loss is validated through comparison with laboratory measure-
ments.

55
Appendix

Part A

Functions used to solve for the wavenumber 0 k0 ,

h i h i
U1 = (1 − eα ) L1 ζ 2L − k2 ν2 / (1 − ν2 ) ; U2 = (1 + eα )ζ L k4 − k41

h i
U3 = (1 − e− β )ζ T kL1 [(1 − 2ν2 ) / (1 − ν2 )] ; U4 = (1 + e− β )k k4 − k41

h i
V1 = (1 − eα )k k2L1 − k2 ; V2 = (1 + eα )2ζ L M1 k

h i h i
−β −β
V3 = (1 − e )ζ T k2L1 −k 2
; V4 = (1 + e )2M1 k 2
− k2L1 / (1 − ν2 )

12E2 (1 − ν2 ) 1 − ν12 E2 1 − ν12


 
L1 = ; M1 =
E1 t31 (1 + ν2 ) (1 − 2ν2 ) 2E1 t1 (1 + ν2 )

α = ζ L λ L ; β = ζ L λT

s s
ω2 ρ 2 (1 + ν2 ) (1 − 2ν2 ) 2ω 2 ρ2 (1 + ν2 )
ζL = k2 − ; ζT = k2 −
E2 (1 − ν2 ) E2

s s
ω2 ρ 2 ω 2 ρ2
λL = ; λT =
E20 − k 2 G2 − k2

56
Overview and Summary

Part B

Principle of Virtual work

The principle of virtual work states that the sum of the works done by all the elements
in one period of the system must do no virtual work when moved through any one of
the virtual displacements [28]

δWi = δαi,m e− j(k x +2mπ/L) x . (8.1)

The total virtual work is the sum of virtual works of the two panels elements, the
virtual work of the translational and rotational springs and from the lumped mass of
the studs. The principle of virtual work requires that

δΠ p1 + δΠ p2 + δΠt1 + δΠt2 + δΠr1 + δΠr2 + δΠm1 + δΠm2 = 0. (8.2)

Virtual work of panel elements: Isotropic

The virtual work contributed by the two panel elements in one period of a panel ele-
ment can expressed as

ˆL 
∂4 W1 ∂2 W1

δΠ p1 = D1 + m p1 2 − jωρo Φ1 + jωρc Φ2 δW1∗ dx at y = 0, (8.3)
∂x4 ∂t
x =0

ˆL 
∂4 W2 ∂2 W2

δΠ p2 = D2 + m p2 2 − jωρc Φ2 + jωρo Φ3 δW2∗ dx at y = H. (8.4)
∂x4 ∂t
x =0

where δWi∗ is the complex conjugate of the virtual displacement.

Virtual work of panel elements: Orthotropic

The virtual work contributed by the two panel elements in one period of a panel ele-
ment can expressed as

57
Sathish Kumar

ˆL 
∂4 W1 ∂4 W1 ∂4 W1 ∂2 W1

δΠ p1 = Dx1 + 2D xz1 + D z1 + m p1 − jωρo Φ1 + jωρc Φ2
∂x4 ∂x2 ∂z2 ∂z4 ∂t2
x =0
δW1∗ dx at y = 0,
(8.5)

ˆL 
∂4 W2 ∂4 W2 ∂4 W2 ∂2 W2

δΠ p2 = Dx2 + 2D xz2 + D z2 + m p2 − jωρc Φ2 + jωρo Φ3
∂x4 ∂x2 ∂z2 ∂z4 ∂t2
x =0
δW2∗ dx at y = H.
(8.6)

where δWi∗ is the complex conjugate of the virtual displacement.

Virtual work of translational springs


The contribution to the virtual work by the translational spring is equal to

+∞ +∞
" #

δΠt1 = Kt (W1 (0) − W2 (0)) δα1,m = Kt ∑ α1,n − ∑ ∗
α2,n δα1,m , (8.7)
n=−∞ n=−∞

+∞ +∞
" #
δΠt2 = ∗
Kt (W2 (0) − W1 (0)) δα2,m = Kt ∑ α2,n − ∑ ∗
α1,n δα2,m . (8.8)
n=−∞ n=−∞

Virtual work of rotational springs


The contribution to the virtual work by the rotational spring is equal to

 0
  0 2mπ


δΠr1 = Kr W1 (0) − W2 (0) j k x + δα1,m
L
+∞ +∞
"    #  
2nπ 2nπ 2mπ
= Kr ∑ α1,n k x + − ∑ α2,n k x + kx + ∗
δα1,m ,
n=−∞ L n=−∞ L L
(8.9)

58
Overview and Summary

 0 0
  2mπ


δΠr2 = Kr W2 (0) − W1 (0) j k x + δα2,m
L
+∞ +∞
"    #  
2nπ 2nπ 2mπ
= Kr ∑ α2,n k x + − ∑ α1,n k x + kx + ∗
δα2,m .
n=−∞ L n=−∞ L L
(8.10)

Virtual work of lumped masses


The contribution to the virtual work by the lumped mass is equal to

+∞
" #

δΠm1 = −ω 2 MW1 (0) δα1,m = −ω 2 M ∑ ∗
α1,n δα1,m , (8.11)
n=−∞

+∞
" #
δΠm2 = −ω 2 ∗
MW2 (0) δα2,m 2
= −ω M ∑ ∗
α2,n δα2,m . (8.12)
n=−∞

59
Bibliography

[1] Jean F. Allard and Yvan Champoux. New empirical equations for sound propa-
gation in rigid frame fibrous materials. Journal of the Acoustical Society of America,
91:3346–3353, 1992.

[2] Leo L. Beranek. Noise and Vibration Control Engineering. John Wiley & Sons, 1992.

[3] U. Carlsson. Acoustical and vibrational aspects of future railway vehicles. Tech-
nical report, KTH-FKT, 1997.

[4] S. I. Cha and H. H. Chun. Insertion loss prediction of floating floors used in ship
cabins. Applied Acoustics, 69:913–917, 2008.

[5] R. J. M. Craik. Non-resonant sound transmission through double walls using


statistical energy analysis. Applied Acoustics, 64:325–341, 2003.

[6] R. J. M. Craik and R. S. Smith. Sound transmission through double leaf


lightweight partitions part i: airborne sound. Applied Acoustics, 61:223 – 245, 2000.

[7] R. J. M. Craik and R. S. Smith. Sound transmission through lightweight parallel


plates. part ii: structure-borne sound. Applied Acoustics, 61:247 – 269, 2000.

[8] L. Cremer, M. Heckl, and B. A. T. Petersson. Structure-Borne Sound. Springer, 2005.

[9] C. L. Dym and M. A. Lang. Transmission of sound through sandwich panels.


Journal of the Acoustical Society of America, 56:1523–1532, 1974.

[10] C. L. Dym and C. S. Ventres. Transmission of sound through sandwich panels: A


reconsideration. Journal of the Acoustical Society of America, 59:364–367, 1976.

[11] Frank J. Fahy. Sound and Structural Vibration: Radiation, Transmission and Response.
Academic Press, 1987.

[12] L. Feng. Acoustical Measurements. KTH-AVE Lecture notes, 2007.

61
Sathish Kumar

[13] L. Feng and S. Kumar. On application of radiation loss factor in the prediction of
sound transmission loss of a honeycomb panel. International Journal of Acoustics
and Vibration, 17:47–51, 2012.

[14] S. Finnveden and M. Barbagallo. A cavity-wall element for statistical energy anal-
ysis of double walls. Technical report, KTH-AVE, 2012.

[15] R. D. Ford, P. Lord, and A. W. Walker. Sound transmission through sandwich


constructions. Journal of Sound and Vibration, 5:9–21, 1967.

[16] R. E. Halliwell, T. R. T. Nightingale, A. C. C. Warnock, and J. A. Birta. Gypsum


board walls: transmission loss data. Technical report, Institue of Research in Con-
struction, 1998.

[17] M. Heckl. The tenth sir richard fairey memorial lecture: Sound transmission in
buildings. Journal of Sound and Vibration, 77:165–189, 1981.

[18] L. Kari, K. Lindgren, L.Feng, and A.C. Nilsson. Constrained polymer layers to
reduce noise: reality or fiction?- an experimental inquiry into their effectiveness.
Polymer esting, 21:949–958, 2002.

[19] J. E. M. Kerwin. Damping of flexural waves by a constrained viscoelastic layer.


Journal of the Acoustical Society of America, 31:952–962, 1959.

[20] G. Kurtze and B. Watters. New wall design for high transmission loss or high
damping. Journal of the Acoustical Society of America, 31:739–748, 1959.

[21] J. H. Lee and J. Kim. Analysis of sound transmission through periodically stiff-
ened panels by space-harmonic expansion method. Journal of Sound and Vibration,
251:349 – 366, 2002.

[22] J. Legault and N. Atalla. Sound transmission through a double panel structure
periodically coupled with vibration insulators. Journal of Sound and Vibration, 329:
3082 – 3100, 2010.

[23] J. Legault and N. Atalla. Numerical and experimental investigation of the effect of
structural links on the sound transmission of a lightweight double panel structure.
Journal of Sound and Vibration, 324:712 – 732, 2009.

[24] F. G. Leppington, E. G. Broadbent, and K. H. Heron. The acoustic radiation ef-


ficiency of rectangular panels. Proceedings of the Royal Society of London, A382:
245–271, 1982.

62
Overview and Summary

[25] R. H. Lyon and R. G. DeJong. Theory and Application of Statistical Energy Analysis.
Butterwirth-Heinemann, 1995.

[26] de M. J. Regt. Comparison between measured and calculated acoustical properties


of floating floors for use on board ships. In Proceedings of 1981 International Con-
ference on Noise Control Engineering, Amsterdam, Netherlands, pages 705–710, 1981.

[27] G. Maidanik. Response of ribbed panels to reverberant acoustic fields. Journal of


the Acoustical Society of America, 34:809–826, 1962.

[28] D. J. Mead and K. K. Pujara. Space-harmonic analysis of periodically supported


beams: response to convected random loading. Journal of Sound and Vibration, 14:
525 – 532, IN9, 533–541, 1971.

[29] J. A. Moore and R. H. Lyon. Sound transmission loss characteristics of sandwich


panel constructions. Journal of the Acoustical Society of America, 89:777–791, 1991.

[30] K. A. Mulholland, H. D. Parbrook, and A. Cummings. The transmission loss of


double panels. Journal of Sound and Vibration, 6:324 – 334, 1967.

[31] A. C. Nilsson. Wave propagation and sound transmission through sandwich


plates. Journal of Sound and Vibration, 138:73–94, 1990.

[32] A. C. Nilsson. Reduction of structure-borne sound in simple ship structures :


Results of model tests. Journal of Sound and Vibration, 61:45–60, 1978.

[33] A. C. Nilsson. Some acoustical properties of floating-floor constructions. Journal


of the Acoustical Society of America, 61:1533–1539, 1977.

[34] E. Nilsson and A. C. Nilsson. Prediction and measurement of some dynamic prop-
erties of sandwich structures with honeycomb and foam cores. Journal of Sound
and Vibration, 251:409–430, 2001.

[35] A. Pellicier and N. Trompette. A review of analytical methods, based on the wave
approach, to compute partitions transmission loss. Applied Acoustics, 68:1192 –
1212, 2007.

[36] Allan D. Pierce. Acoustics : an introduction to its physical principles and applications.
McGraw-Hill Book Co., New York, 1981.

[37] J. Poblet-Puig, A. Rodriguez-Ferran, C.Guigou-Carter, and M. Villot. The role


of studs in the sound transmission of double walls. Acta Acustica United with
Acustica, 45:555–567, 2009.

63
Sathish Kumar

[38] A. J. Price and M. J. Crocker. Sound transmission through double panels using
statistical energy analysis. Journal of the Acoustical Society of America, 47:683–693,
1970.

[39] J. Wang, T. J. Lu, J. Woodhouse, R. S. Langley, and J. Evans. Sound transmission


through lightweight double-leaf partitions: theoretical modelling. Journal of Sound
and Vibration, 286:817 – 847, 2005.

[40] T. Wang, V. S. Sokolinsky, S. Rajaram, and S. R. Nutt. Assessment of sandwich


models for the prediction of sound transmission loss in unidirectional sandwich
panels. Applied Acoustics, 66:245–262, 2005.

[41] F. X. Xin and T. J. Lu. Transmission loss of orthogonally rib-stiffened double-panel


structures with cavity absorption. Journal of the Acoustical Society of America, 129:
1919–1934, 2011.

[42] D. Zenkert. An Introduction to Sandwich Construction. Engineering Material Advi-


sory services Ltd, London, 1995.

64

You might also like