You are on page 1of 280

Thermoacoustic refrigerators : experiments and scaling

analysis
Citation for published version (APA):
Li, Y. (2011). Thermoacoustic refrigerators : experiments and scaling analysis. Technische Universiteit
Eindhoven. https://doi.org/10.6100/IR716451

DOI:
10.6100/IR716451

Document status and date:


Published: 01/01/2011

Document Version:
Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:


• A submitted manuscript is the version of the article upon submission and before peer-review. There can be
important differences between the submitted version and the official published version of record. People
interested in the research are advised to contact the author for the final version of the publication, or visit the
DOI to the publisher's website.
• The final author version and the galley proof are versions of the publication after peer review.
• The final published version features the final layout of the paper including the volume, issue and page
numbers.
Link to publication

General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners
and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
• You may not further distribute the material or use it for any profit-making activity or commercial gain
• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please
follow below link for the End User Agreement:
www.tue.nl/taverne

Take down policy


If you believe that this document breaches copyright please contact us at:
openaccess@tue.nl
providing details and we will investigate your claim.

Download date: 02. Oct. 2020


Thermoacoustic Refrigerators:
Experiments and Scaling Analysis
Copyright © 2011 by Yan Li, Eindhoven, The Netherlands.
All rights are reserved. No part of this publication may be reproduced, stored in a
retrieval system, or transmitted, in any form or by any means, electronic,
mechanical, photocopying, recording or otherwise, without prior permission of the
author.

Printed by Print Service Technische Universiteit Eindhoven

Cover design by Paul Verspaget

A catalogue record is available from the Eindhoven University of Technology


Library

Li, Yan

Thermoacoustic Refrigerators: Experiments and Scaling Analysis /


by Yan Li.-
Eindhoven: Technische Universiteit Eindhoven, 2011.
Proefschrift.-ISBN 978-90-386-2670-3
NUR 929

This research was financially supported by MicroNed, grant number 1-B-7


Thermoacoustic Refrigerators:
Experiments and Scaling Analysis

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit Eindhoven,
op gezag van de rector magnificus, prof.dr.ir. C.J. van Duijn, voor een commissie
aangewezen door het College voor Promoties in het openbaar te verdedigen op
donderdag 27 oktober 2011 om 14.00 uur

door

Yan Li

geboren te LiaoNing, China


Dit proefschrift is goedgekeurd door de promotoren:

prof.dr.ir. H.J.M. ter Brake


en
prof.dr. A.T.A.M. de Waele

Copromotor:
dr.ir. J.C.H. Zeegers

Ten years living and dead have drawn apart


I do nothing to remember
But I cannot forget
Your lonely grave a thousand miles away
Nowhere can I talk of my sorrow
Even if we met, how would you know me
My face full of dust
My hair like snow
In the dark of night, a dream suddenly, I am home
You by the window
Doing your hair
I look at you and cannot speak
Your face is streaked by endless tears
Year after year must they break my heart
These moonlit nights?
That low pine grave?

By Su Shi

谨以此文赠给我的弟弟
Contents

1 Introduction 1
1.1 Thermoacoustics · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 1
1.2 History of thermoacoustics· · · · · · · · · · · · · · · · · · · · · · · · ·· · ·· · 2
1.3 Objective of present work · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 6
1.4 The scope of this thesis · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 7

2 Basic theory of thermoacoustics 8


2.1 Wave equation and total energy flow · · · · · · · · · · · · · · · · · · · · 8
2.2 Acoustic energy · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 22

3 Standing-wave systems 26
3.1 Introduction · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 26
3.2 Physical description of standing-wave systems · · · · · · · · · · · · · 29
3.3 Modeling standing-wave systems · · · · · · · · · · · · · · · · · · · · · · · 32
3.3.1 Zero viscosity · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 33
3.3.2 General analysis with viscosity included · · · · · · · · · · · 46
3.3.3 General analysis of “TAC” · · · · · · · · · · · · · · · · · · · · · 51
3.4 Experimental results · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 58
3.4.1 Experimental set-up · · · · · · · · · · · · · · · · · · · · · · · · · · · 58
3.4.2 Measurements · · · · · · · · · · · · · · · · · · ·· · · · · · · · · · · · · 66
3.4.3 Theoretical computation · · · · · · · · · · ··· · · · · · · · · · · · 76
3.4.4 Conclusions· · · · · · · · · · · · · · · · · ·· · · · ·· · · · · · · · · · · · 84

4 Traveling-wave systems 85
4.1 Introduction · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 85
ii Contents

4.2 Physical description of traveling-wave systems · · · · · · · · · · · · 86


4.3 Modeling traveling-wave systems · · · · · · · · · · · · · · · · · · · · · · · 88
4.4 Optimizing regenerator material· · · · · · · · · · · · · · · · · · · · · ·· · · 114
4.4.1 Introduction· · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 114
4.4.2 Regenerator materials· · · · · · · · · · · · · · · · · · · · · · · · · · 114
4.4.3 Selection criteria· · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·· 120
4.4.4 Experimental set-up: a coaxial traveling-wave engine· 125
4.4.5 Energy balance in the experimental set-up· · · · · · · · ·· · 129
4.4.6 Measurement equipment and data handling· · · · · · · · ·· 130
4.4.7 Measuring procedure· · · · · · · · · · · · · · · · · · · · · · · · · · · 133
4.4.8 Results and discussion· · · · · · · · · · · · · · · · · · · · · · · · · · 135
4.4.9 Conclusion· · · · · · · · · · · · · · · · · · · · · · · · · · · ··· · · · ·· · 149
4.5 Experiments on a thermoacoustic refrigerator · · · · · · · · · · · · · 151
4.5.1 Introduction· · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 151
4.5.2 Design considerations· · · · · · · · · · · · · · · · · · · · · · · · · · 151
4.5.3 Experimental set-up· · · · · · · · · · · · · · · · · · · · · · · · · · · 152
4.5.4 Measurement equipment and data handling· · · · · · · · ·· 157
4.5.5 Losses· · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·· · · · · · · 160
4.5.6 Measuring procedure· · · · · · · · · · · · · · · · · · · · · · · · · · · 162
4.5.7 Results and discussion· · · · · · · · · · · · · · · · · · · · · · · · · · 163
4.5.8 Discussions· · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·· 183
4.5.9 Conclusion· · · · · · · · · · · · · · · · · · · · · · · · · · · ··· · · · ·· ·· 185

5 Scaling considerations 186


5.1 Introduction· · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·· · · · · · · · 186
5.2 Standing-wave systems · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 187
5.2.1 Constant temperature difference over the stack · · · · · 187
5.2.2 Constant time-averaged total energy flow · · · · · · · · · · 199
5.2.3 Constant time-averaged total energy flow density · · · · 201
5.3 Traveling-wave systems · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·· 202
5.4 Conclusions · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · 210
6 Conclusions and recommendations 211
6.1 Conclusions 211
Contents iii

6.2 Recommendations 214

Appendices 215
A Momentum equations derivation· · · · · · · · · · · · · · · · · · · · · · ·· · · · · · · 215
B Derivation of the temperature of the solid plate· · · · · · · · · · · · · · · · · · 217
C Derivation of the temperature oscillation of the fluid layer· · · · · · · · · 218
D Derivation of the time-averaged total energy flow· · · · · · · · · · · · · · · · 222
E Derivation of the decoupling the sound field into standing-wave and
traveling-wave components· · · · · · · · · · · · · · · · · ·· · ·· · · · · · · · · · · · ·· 224
F Computation of loop section in a traveling-wave system· ·· ·· · · · · · ·· 226
G Transmission of acoustic impedance of a uniform pipe· · ·· · ·· · ·· · ·· 235
H Fortran code for computation of traveling-wave engine· · · ·· · ·· · · · · 237
I The design of the ambient heat exchanger in the traveling-wave
refrigerator· · · · · · · · · · · · · · · · · · · · · ·· · · · ·· · · · · · · · · · · · · · · · · · · · 242
J Time evolution of two orientations: upward and downward in
traveling-wave refrigerator measurement· · · · · · · · · · · · · · · · · · · · · ·· 246
K Acoustic field in the scaled-down standing-wave systems· · · · · · · · · 248

Nomenclature 254
Bibliography 258
Summary 264
Samenvatting 266
Dankwoord 269
Curriculum Vitae 271
Chapter 1

Introduction

1.1 Thermoacoustics

Thermoacoustics is a subject which focuses on the interaction between solid walls


and oscillating fluids from thermodynamic point of view. Under normal conditions,
the periodical adiabatic compression and expansion make the temperature of the
sound propagating medium oscillate with a tiny amplitude, which is hardly
perceptible to human beings. For instance, a normal conversation, scaled as 60 dB,
can produce an excess temperature of only a few hundredths of a degree Celsius. It
is this almost invisibility of the thermodynamic effect which prevented
thermoacoustics from being explored earlier. In 1980, Nikolaus Rott [1] first
introduced the term “thermoacoustics” in a review of his previous work on a theory
for this phenomenon. The theory, known as linear thermoacoustic theory, became
the solid basis of nowadays thermoacoustic investigations and applications. In
recent decades, investigations on the fundamental nature of the problems
encountered in various thermoacoustic devices and explorations on industrial and
household applications are widely carried out in many research groups world-wide.
Many thermoacoustic devices were built and utilized. The performances and
efficiencies are much enhanced. Thermoacoustic devices have the advantages over
the conventional heat pumps and engines, that they have no mechanical moving
parts, which brings high reliability and virtually maintenance-free to the customers.
Moreover, they are environment friendly by using chemically inert working gases.
It is a charming technology for today’s world, which is suffering all sorts of
environmental problems: global warming, ozone depletion and others.
2 Chapter 1

1.2 History of thermoacoustics

The first reported observation about the thermoacoustic effect in the community of
physicists was in the year of 1802 by Bryan Higgins [2]. In 1777, about 10 years
after the discovery of hydrogen, Higgins demonstrated that the burning of
hydrogen produces water. He lowered a vertical glass tube, which was sealed at the
far end, over the flame. What ensued is unexpected “singing”. Later, he also tried
different glass tubes and produced “several sweet tones, according to the width,
length and thickness of the glass jar or sealed tube”. This singing flame aroused the
interest of many investigators. Many explanations for this interesting effect were
proposed, but they were largely incorrect. Jones made a good discussion on all
these theories and modified the Rayleigh’s theory [3]. Later, Putnam and Dennis
gave a wide survey on all sorts of combustion oscillations related to this “singing
flame” [4].
Another interesting thermoacoustical oscillation, namely the Rijke tube, was
reported by Rijke in 1859 [5]. Rijke found that strong oscillations occurred when a
heated wire screen was placed in the lower half of a vertical pipe with two open
ends, as shown in Fig. 1.2.1a. It was also found that the oscillations would stop if
the top of the pipe was closed, implying that the convective air current was
necessary for this phenomenon. Oscillations became strongest when the heated
screen was located one-fourth of the length of the pipe from the bottom end.
Although Rijke gave some explanation, it was thought as inadequate to explain the
detailed heat exchange mechanism causing the oscillations.

Sound generated
Bulb

Tube stem

Heated screen Heat addition


Sound generated
Convection flow
(b)
(a)
Figure 1.2.1: (a) Rijke tube (b) Sondhauss tube.
Introduction 3

In the consequential years, many theoretical analysis and experimental work were
provided to explain this phenomenon qualitatively and quantitatively as well.
Feldman reviewed the literature [6].
The Rijke oscillations are observed in many industrial facilities, like gas furnaces,
oil burners, gas-heated deep fat fryers, and rocket combustion chambers. This
annoying, sometimes even destructive effect, are described as “screaming”,
“screeching”, and “chugging”. An important difference between the Rijke effect
and thermoacoustics is that in a Rijke tube an average velocity is present on top of
the acoustic oscillations.
As mentioned at the beginning in this section, before scientists worked on these
thermoacoustical phenomena, the glass blowers had heard a lot of “glass singing”,
when they blew bulbs on the ends of narrow tubes. Sondhauss was the first to study
experimentally these “singing glasses”. He published his investigations [7] in 1850
on a tube which was open on one end and terminated in a bulb on the other end,
with a steady gas flame applied to the closed bulb-end, as shown in Fig. 1.2.1b.
Such a tube was therefore named as “Sondhauss tube”, which approximates best
what we define today as thermoacoustic oscillations. Sondhauss discovered that a
steady gas flame, applied to the closed bulb end, caused the air in the entire tube to
oscillate and produce a clear sound which was characteristic of the dimension of
the tube. He also observed that larger bulbs and longer tubes produced lower
frequency sounds and that hotter flames produced more intense sounds. Knipp also
observed that thermoacoustic oscillations occurred when a glass vapor trap was
heated and suggested the apparatus could be used as a standard source of sound [8].
The first referred Sondhauss oscillation taking place in the cryogenic research is
“Taconis oscillations”. Taconis observed spontaneous oscillations in a hollow tube
with the upper end closed at room temperature and the lower end immersed in the
liquid helium [9]. He explained how the large thermal gradient along the tube
caused the oscillations. The Taconis oscillations have been investigated
experimentally by Yazaki et al. [10]
In 1878, Lord Rayleigh proposed his criterion on these related thermoacoustical
oscillation phenomena [11]:
“If heat be given to the air at the moment of greatest condensation or be taken from
it at the moment of greatest rarefaction, the vibration is encouraged”.
This qualitative explanation was proved to agree well with extensive experimental
observations and widely accepted by thermoacoustic community.
An important progress came in 1962, when Carter et al. experimentally
investigated the Sondhauss oscillation to determine the feasibility of using the
phenomenon to generate electricity [12]. They found that inserting a bundle of
small glass capillaries at a suitable position inside the Sondhauss tube could greatly
4 Chapter 1

improve the performance. This bundle of small glass capillaries is the so-called
“stack” in modern thermoacoustics. This discovery made the later applications,
using thermoacoustic phenomena, feasible and practical. Extensive studies
following this idea were performed by Feldman in his PhD work [13]. He also
made a review on literature work about Sondhauss tube [14].
Compared with the history of heat-driven oscillations, which is rich and old, the
reverse thermodynamic process, of generation of a temperature gradient by
imposing acoustic oscillations is rather recent. The first work on thermoacoustic
type cooling was carried out by Gifford and Longsworth in 1964 [15]. They
invented the pulse-tube refrigerator driven by a low frequency acoustic wave, to
cool down to a temperature of 150K. Due to the efforts of many researchers, the
pulse-tube has become one of the most favored technologies for cryocooling. A
complete history and review of pulse tube works is given by Radebaugh [16, 17].
More information about modeling and numerical analysis of pulse-tube
refrigerators can be found in [18-19]. In 1975, P. Merkli and H. Thomann reported
their observation of thermoacoustic effects in a resonance tube [20]. They found
cooling in the section of the tube with maximum velocity amplitude and marked
heating in the region of the velocity nodes. They also developed a theoretical
model which agreed with experiment at low amplitudes.
Although much experimental work had been done and after Rayleigh’s qualitative
explanation, researchers got progress in theoretical exploration to quantitatively
describe these thermoacoustic oscillations at a much later time.
The formal study on the theoretical aspect was started by Kramers in 1949 [21]. He
developed a theoretical model to explain “Taconis oscillations”, by employing the
method of solution used previously by Kirchhoff to achieve an exact solution for
gas vibrations in a tube of constant temperature throughout [22]. By confining the
phenomenon to small amplitude wave, he could linearize hydrodynamic equations
of mass, momentum, and energy. Although he successfully separated the wave
components and solved the resulting linearized equations, he was unable to account
for the spontaneous vibrations which were often observed in experiments. He
attributed this unsatisfactory feature of his theory to some neglected terms in
linearizing which were probably not negligible.
Trilling did theoretical analysis on an induced sound field by applying a sudden
temperature variation on the rest boundary of a viscous heat-conducting gas [23].
In his analysis, the temperature at the closed end of a semi-infinite gas-filled pipe
suddenly raised, the gas near the hot wall expanded and moved outwards, function
like a piston. He showed that the magnitude of the pressure pulse generated was
proportional to that of the temperature increase and inversely proportional to the
one-fourth root of the distance traveled.
Introduction 5

Chu published four theoretical papers about heat-generated pressure waves. In the
first paper a modified wave equation with the heat addition as source term was
derived to describe the pressure field generated by a moderate rate of heat release
[24]. In the second paper, Chu analyzed the stability of systems containing a heat
source [25]. In the third paper, Chu and Ying theoretically investigated non-linear
oscillations produced by a sinusoidal heat release from a plane heater located at the
midsection of a completely closed pipe [26]. In the fourth paper, he theoretically
studied a self-sustained, thermally driven, non-linear oscillation in a closed pipe
[14].
The breakthrough came in 1969 by a series of articles of Rott [27-32, 1]. Based on
review of previous works, Rott re-examined the simplifying assumptions used in
Kramers’ work and abandoned incorrect ones. His remarkable work has built a
solid theoretical basis of thermoacoustics, and becomes one of the most refered
papers in modern thermoacoustics. The review article, published by Rott in 1980
[1] on summary of his previous results, inaugurated an active and prolific era in
thermoacoustics. Enormous related projects have been conducted and progresses
have been achieved. The Condensed Matter and Thermal Physics group of Los
Alamos National Laboratory started a research program to apply Rott’s theory to
build functional devices. In 1988, G. Swift published a comprehensive article
addressing important aspects of thermoacoustic devices [33]. In 2000, S. Backhaus
and G. Swift [34] presented an efficient thermoacoustic traveling-wave engine
which made this novel thermoacoustic technology competitive with present
conventional thermal machines widely used commercially. This new technology
has been now investigated and efforts have been put to applications in industrial
and normal household facilities, in a world-wide scale: the US, Canada, France,
Mexico, the Netherlands, China, Japan, and other countries. In the last few
decades, thermoacoustics has gone through a prosperous time. Much progress and
achievement have been collected and reviewed by Garrett [35].
6 Chapter 1

1.3 Objective of present work

Section 1.2 describes the history of thermoacoustics, and from the developments of
the work at Los Alamos by Swift and coworkers, the work at Penn State University
by Garrett and coworkers, as well as developments at ECN in the Netherlands and
many other laboratories in the world, traveling wave thermoacoustic engines are
built, and sized often as large apparatus, having a length of 3 meters up till sizes of
even 25 meters long. Also standing wave devices are generally 50 cm or longer.
The apparatus that has been built at Penn State University in the group of Steven
Garrett, although very compact is also of a size on the order of 0.50 m long and
0.25 m diameter.
In space application as well as in laptop computers or even mobile phones there is
a strong need of cooling devices to cool away the heat that is generated by the ultra
fine IC components that have a high intensity local heat production. That motivated
also this project out of the perspective of the MicroNed grant, where the focus is on
cooling of small scale (space) devices. Now from thermoacoustic point of view,
small scales will mean that high frequencies have to be used. Using high
frequencies on the one hand pushes up the criteria on downsizing all the
components that are needed to build such a device. But apart from that
thermoacoustic heat transport is strongly related to temperature differences over a
stack or regenerator of finite length. When downsizing the length it will mean that
thermal gradients will increase and there must be a limit on scalability of such
devices due to the laws of thermodynamics i.e. heat conduction or any other loss
processes. This issue, the rules for scaling down thermoacoustic refrigerators to
miniature size, and to discover the limitation of that is one of the main topics of
this dissertation. The main goal is to provide some guidance for the design of
small-scale thermoacoustic machines. The two types of thermoacoustic
refrigerators, standing-wave and traveling-wave, are both investigated for scaling.
The basis of this scaling forms an investigation by means of analysis using the
thermoacoustic equations, and applying them to both types of devices. Cooling
rates, heat conduction, and power production are investigated analytically and
scaling rules can be derived to study the influence of scaling. Apart from that the
modelling results are partially verified by comparing them with experimental
apparatus as built by Swift, as well as in our own laboratory. It gives this work a
solid foundation for future design work on scaling of thermoacoustic refrigerators.
Introduction 7

1.4 The scope of this thesis


This thesis presents the following contents: Chapter 2 is dedicated to a brief review
of the basic theory of thermoacoustics, which is the widely used linear
thermoacoustic theory developed by Rott and implemented by Swift. Chapter 3 is
concerned with standing-wave refrigerator systems. The working principles to
generate cooling by an acoustic wave are described. The analytic expressions are
applied into a model, that describes a 25 cm tubular standing wave resonator. In
chapter 4 the theory of the traveling wave systems is derived and analytical
expressions are found for the thermoacoustic equations describing the energy flows
in this system. By applying these expressions into a numerical model a fast
numerical design tool in Fortran has been written by which traveling wave systems
can be studied efficiently. This model is applied to Swifts traveling wave engine
described in reference [34] and shows good agreement with DeltaE computations.
Apart from that chapter 4 describes two experiments with traveling wave
apparatus, one co-axial type as developed by ECN, and a new concept namely a
motor driven 1.3 meter long tubular traveling wave cooler, developed at TU/e. The
results of the measurements are compared with the model to obtain insight
concerning the validation. An important result that should already be mentioned
here is that our numerical design tool for the traveling wave system indicated that
for these smaller scale systems it is not necessary to contain a compliance to build
a cooler. A best performance can be obtained with a single size diameter feedback
tube. The experimental system of the TU/e cools very well by using such one
diameter size feedback tube. Finally in chapter 5 the results of chapters 3 and 4 are
combined into two analytical scaling models for standing wave as well as traveling
wave systems. These models that start from a macroscopic known apparatus
demonstrate that scaling towards millimeter size devices leads to a strong decrease
of the performance. Chapter 6 concludes this thesis. In the appendixes some
detailed experimental data and some mathematical derivations are presented.
Chapter 2

Basic theory of thermoacoustics

2.1 Wave equation and total energy flow

Geometry
The widely-used linear thermoacoustic theory as known today was first developed
by Rott and reviewed by Swift [33]. First, the linearization of the Navier-Stokes
and continuity equations gives us the wave equation of thermoacoustics. Next,
energy conservation and heat transfer equations provide us the total energy flow
expression.
As shown in Figs. 2.1.1 and 2.1.2, we consider a stack of parallel plates in an
acoustic field. The conditions are defined in Fig.2.1.2. The x axis is along the
direction of sound propagation, the y axis normal to the fluid-solid boundary. y=0
is located in the center of the fluid. The thickness of the fluid layer between two
adjacent stack plates is 2 y0 as shown in Fig. 2.1.2. The y′ axis for the solid is
normal to the fluid-solid boundary, with y′ = 0 in the center of the solid and
y′ = l at the boundary, see Fig. 2.1.2. Axes y and y′ have opposite directions.

Sound wave

x Stack of plates

Figure 2.1.1: Geometry used for a multi-plate thermoacoustic system.


Basic theory of thermoacoustics 9

Fluid
Solid

Fluid
y′ x 2l
Solid
y
2 y0
x
Fluid
Solid

Figure 2.1.2: Geometry used for multi-plate stack.

Basic Equations
Assume that all variables oscillate at a single angular frequency ω and use an
expansion up to first-order in the acoustic amplitude for all variables.
p = pm + Re[p1 ( x )eiω t ] (2.1.1)
ρ = ρ m ( x ) + Re[ ρ1 ( x , y ) e iω t ] (2.1.2)
r v v
V = x Re[u1 ( x, y)eiω t ] + y Re[v1 ( x, y)eiω t ] (2.1.3)
T = Tm ( x ) + Re[ T1 ( x , y ) e iω t ] (2.1.4)
iω t
Ts = Tm ( x ) + Re[ Ts1 ( x , y ′) e ] (2.1.5)
s = sm ( x) + Re[s1 ( x, y )eiω t ] (2.1.6)
The subscripts “m” indicates mean value and “s” indicates solid. Throughout this
study, complex quantities are represented by boldface type, with exceptions:
I) the definition i = − 1 and
II) The Rott’s functions: f v , fκ and ε s .
Thus, variables like p 1 ( x), ρ1 ( x) and etc. are complex amplitudes.
Note that we also make the following assumptions:
1. The theory is linear, zero-order and first-order terms are kept for equations
other than energy equations. For energy equations, second–order terms are
considered.
r
2. The zero order average fluid velocity Vm = 0
3. The solid is perfectly rigid.
4. The working fluid is considered to be an ideal gas.
5. Gravity is neglected.
6. Second viscosity is neglected [37].
10 Chapter 2

7. The pressure dependency of the viscosity and thermal conductivity is


neglected.
In the geometry of Fig 2.1.2, we begin with deriving an expression for the x
component of the fluid velocity. The momentum equation of a compressible,
viscous fluid in two dimensions [36],
 ∂u ∂u ∂u  ∂p ∂   ∂u 2  ∂u ∂v  
ρ  +u + v  = − + µ 2 −  +  
 ∂t ∂x ∂y  ∂x ∂x   ∂x 3  ∂x ∂y  
∂   ∂u ∂v 
+ µ  +  , (2.1.7)
∂y   ∂y ∂x 
where µ is the dynamic viscosity. The viscosity µ is a function of temperature.
Rearranging the terms in Eq. (2.1.7) yields:
 ∂u ∂u ∂u  ∂p  ∂ 2u ∂ 2u  µ ∂  ∂u ∂v 
ρ  +u + v  = − + µ  2 + 2  +  + 
 ∂t ∂x ∂y  ∂x  ∂x ∂y  3 ∂x  ∂x ∂y 
4 ∂µ ∂u 2 ∂µ ∂v ∂µ ∂u ∂µ ∂v
+ − + + . (2.1.8)
3 ∂x ∂x 3 ∂x ∂y ∂y ∂y ∂y ∂x
Substitute the variables expressed from Eq. (2.1.1) to Eq. (2.1.6), and keep the
terms till first-order:
dp1  ∂ 2u1 ∂ 2u1  µ  ∂ 2u1 ∂ 2 v1 
iωρ mu1 = − + µ  2 + 2  +  2 + 
dx  ∂ x ∂ y  3  ∂ x ∂x ∂ y 
4 ∂µ ∂u1 2 ∂µ ∂v1 ∂µ ∂u1 ∂µ ∂v1
+ − + + . (2.1.9)
3 ∂x ∂x 3 ∂x ∂y ∂y ∂y ∂y ∂x
Since
∂µ dµ ∂T dµ ∂T1
= = , (2.1.10)
∂y dT ∂y dT ∂y
∂µ ∂u1
Eq. (2.1.10) leads to the conclusion that the two terms in Eq. (2.1.9) and
∂y ∂y
∂µ ∂v 1
are negligible second-order terms. So, now, Eq. (2.1.9) can be reduced to:
∂y ∂x
dp1  ∂ 2u1 ∂ 2u1  µ  ∂ 2u1 ∂ 2 v1 
iωρ mu1 = − + µ  2 + 2  +  2 + 
dx  ∂x ∂y  3  ∂x ∂x∂y 
4 ∂µ ∂u1 2 ∂µ ∂v1
+ − . (2.1.11)
3 ∂x ∂x 3 ∂x ∂y
Similar to Eq. (2.1.10),
∂µ dµ ∂T dµ  dTm ∂T1 
= =  +  (2.1.12)
∂x dT ∂x dT  dx ∂x 
Basic theory of thermoacoustics 11

The substitution of Eq. (2.1.12) in Eq. (2.1.11) and neglecting second-order


variation terms, Eq. (2.1.11) can be finally expressed as:
dp1  ∂ 2u ∂ 2u  µ  ∂ 2u ∂ 2 v1 
iωρ mu1 = − + µ  21 + 21  +  21 + 
dx  ∂x ∂y  3  ∂x ∂x∂y 
4 dµ dTm ∂u1 2 dµ dTm ∂v1
+ − . (2.1.13)
3 dT dx ∂x 3 dT dx ∂y
The variations related to x, which is the axial wave propagation direction, are of the
order of the radian wavelength D = λ / 2π , where λ is the wavelength. Those in
the perpendicular y direction relate to the viscous penetration depth. Therefore, we
know that u1 v1 is of the order of D δ v , ∂ ∂x is of the order of 1 D , ∂ ∂y is of
the order of 1 δ v . Here,
δ v = 2 µ ( ρω ) (2.1.14)
∂ u1
2
∂ 2 v1
is the viscous penetration depth. Since δ v << D , the terms µ and µ can
∂x 2 ∂x∂y
be neglected compared with µ ∂ 2 u 1 ∂y 2 .
Therefore, Eq. (2.1.13) reduces to
dp1 ∂ 2u 4 dµ dTm ∂u1 2 dµ dTm ∂v1
iωρ mu1 = − + µ 21 + − . (2.1.15)
dx ∂y 3 dT dx ∂x 3 dT dx ∂y
In normal working conditions, the temperature gradient is not extremely large. In
general, the viscosity can be approximately described as [38]:
µ = µ 0 (T / T0 ) µ
b
(2.1.16)
For T0=300 K, values µ 0 and bµ for some gases, of common interest in
thermoacoustics, are listed in the table 1.1.I.

bµ T0=300 K
T 
µ = µ 0  
µ 0 (kg/m·s) bµ
 T0 
air 1.85E-5 0.76
nitrogen 1.82E-5 0.69
helium 1.99E-5 0.68
neon 3.2E-5 0.66
argon 2.3E-5 0.85
xenon 2.4E-5 0.85
Table 1.1.I Approximate values µ 0 and bµ for some gases
Therefore, the ratios of terms in Eq. (2.1.15) are
12 Chapter 2

∂ 2u1 ∂ 2u1
µ T
∂y 2 3 ∂y 2
= (2.1.17)
4 dµ dTm ∂u1 4bµ dTm ∂u1
3 dT dx ∂x dx ∂x
2
D
is of order >>1, and
δν 2
∂ 2u1 ∂ 2u1
µ T
∂y 2 3 ∂y 2
= (2.1.18)
2 dµ dTm ∂v1 2bµ dTm ∂v1
3 dT dx ∂y dx ∂y
D2
is of order >>1.
δν 2
So, for widely used working gases, in normal working conditions, (without
extremely large temperature gradient dTm / dx ), the last two terms in Eq. (2.1.15)
can be neglected.
Therefore, the momentum equation can be reduced to:
dp 1 ∂ 2 u1
iωρ m u1 = − +µ . (2.1.19)
dx ∂y 2
This is the description of the oscillatory velocity profile as dependant on the
oscillatory pressure gradient including viscous terms.
With boundary conditions: at y = 0 , because of the symmetry, ∂u1 / ∂y = 0 , and
at y = y 0 , because of the solid wall, u1 = 0 , the solution of (2.1.19) follows (see
Appendix A)
i dp1  cosh[(1 + i ) y δ v ] 
u1 = 1 − .
ωρ m dx  cosh[(1 + i ) y 0 δ v ] 
(2.1.20)

As an illustration of the velocity profile, an example is plotted based on Eq. (2.1.20)


in Fig 2.1.3. Here, the velocity variation along the y direction is shown with time as
a parameter. In this example, the following parameters are adopted:
1. standing wave field inside the resonator tube: p1 = p A sin( x / D )
2. p A = 0.1 bar, for helium in 300K and 1 bar of mean pressure.
3. y0 = 2.0 × δν , δν = 2µ / ρ mω is the viscous penetration depth of the fluid at
(1000Hz, 1 bar, and 300K for helium).
4. The computed position is at the middle point of the resonator tube, i.e. x=λ/8,
where λ is the wavelength. The resonator tube length is λ/4.
Basic theory of thermoacoustics 13

ωt=0*(π/4)
ωt=1*(π/4)
70 ωt=2*(π/4)
60
ωt=3*(π/4)
ωt=4*(π/4)
50
ωt=5*(π/4)
40 ωt=6*(π/4)
30 ωt=7*(π/4)
20
10
u1 (m/s)

0
-10 0.0 0.5 1.0 1.5 2.0

-20 y/δν
-30
-40
-50
-60
-70

Figure 2.1.3: Relation between velocity in the x direction and position


perpendicular to that (y direction) at different moments in time.

After deriving the axial flow velocity from the momentum equation, we now
consider the temperature of the solid plate Ts ( x, y , t ) . The following equation
holds:
∂Ts
= κ s ∇ 2Ts , (2.1.21)
∂t
where κ s = K s ρ s cs is the thermal diffusivity of the solid, and K s , ρ s , cs are the
thermal conductivity, density, and specific heat per unit mass, respectively. The
solid’s thermal diffusivity is considered as constant.
Substitution of Eq.(2.1.5) into Eq. (2.1.21) to first order yields:
 d 2T ∂ 2Ts1 iω t ∂ 2Ts1 iω t 
Ts1 ⋅ eiω t ⋅ iω = κ s  2m + e + e . (2.1.22)
 dx ∂x 2 ∂y′2 
Similar to the reduction of Eq. (2.1.13) and (2.1.15), it can be seen that
(∂ 2
)( )
Ts1 ∂x 2 / ∂ 2 Ts1 ∂y ′ 2 ~ (δ s / D ) 2 << 1 , where δ s = 2κ s / ω is the solid’s
thermal penetration depth, and
(d 2Tm / dx 2 ) /(∂ 2 Ts1 / ∂y ′ 2 ) ~ (δ s / D ) 2 /(Ts1 / Tm ) << 1 .
Thus, Eq. (2.1.22) reduces to
∂ 2 Ts1
iωTs1 = κ s . (2.1.23)
∂y ′ 2
The temperature of the plate can be derived from this equation as (see Appendix B)
14 Chapter 2

cosh[(1 + i ) y ′ δ s ]
Ts1 = Tb1 .
cosh[(1 + i )l δ s ]
(2.1.24)

where Tb1 is temperature amplitude at the boundary, and is given by Eq. (C.22) in
appendix C.

The temperature in the fluid is found from the general equation of heat transfer [36].
 ∂s v v  r
+ V ⋅ ∇s  = ∇ ⋅ (K∇T ) + (terms quadratic in velocity).
v
ρT  (2.1.25)
 ∂t 
Here, s is the fluid entropy per unit mass.
From thermodynamics, it is known that
( )
ds = c p / T dT − (β / ρ )dp , (2.1.26)
where β is its isobaric thermal expansion coefficient and equal to 1/T for ideal gas.
Substitution of Eq. (2.1.1) to (2.1.6) into (2.1.25), using Eq. (2.1.26) and keeping
the first order terms, Eq. (2.1.25) becomes
2
dT dK  dTm  d 2Tm
ρ m c p T1iω ⋅ e iω t
− p1iω ⋅ e iω t
+ ρ mu1c p m eiω t =   +K
dx dT  dx  dx 2
dK dTm ∂T1 iω t dK dTm ∂T1 iω t ∂ 2T1 iω t ∂ 2T1 iω t
+ e + e +K 2 e +K 2 e . (2.1.27)
dT dx ∂x dT dx ∂y ∂x ∂y
Compare the terms on the right-hand side of Eq. (2.1.27) with the very last one,
∂ 2T  δ 
2 2
dK  dTm   T1 
  K 21 ~  κ    <<1 (2.1.28a)
dT  dx  ∂y D  T 
∂ 2 T1  δ κ 
2
d 2Tm  T1 
K K ~    <<1 (2.1.28b)
dx 2 ∂y 2  D  T 
dK dTm ∂T1 ∂ 2T  δ 
2

K 21 ~  κ  <<1 (2.1.28c)
dT dx ∂x ∂y D 
dK dTm ∂T1 ∂T δ
2
K 21 ~ κ <<1. (2.1.28d)
dT dx ∂y ∂y D
Thus, neglecting the relatively small terms compared in Eq. (2.1.28a) to (2.1.28d),
Eq. (2.1.27) reduces to
 dTm  ∂ 2T1
ρ m c p  iωT1 + u1  − iωp1 = K 2 . (2.1.29)
 dx  ∂y
Solving this second order differential equation, the temperature oscillation in the
fluid layer can be obtained as (see Appendix C)
p1 1  σ cosh[(1 + i ) y / δ v ]  dp1 dTm
T1 = − 1 − 
ρ mc p ρ mω  (σ − 1)cosh[(1 + i ) y0 / δ v ]  dx dx
2 
Basic theory of thermoacoustics 15

 p
− 1 +
(dp1 dx )(dTm / dx ) 1 + ε s f v  cosh[(1 + i )y / δ κ ] ,
(σ − 1)ρ mω 2  f k  (1 + ε s )cosh[(1 + i ) y0 / δ κ ]
(2.1.30)
 ρ m c p
where
σ = c p µ / K = δ v2 / δ κ2 (2.1.31)
is the Prandtl number, and the Rott’s functions
tanh[(1 + i ) y 0 / δν ]
fν = (2.1.32)
(1 + i ) y 0 / δν
tanh[(1 + i ) y 0 / δ κ ]
fκ = (2.1.33)
(1 + i ) y 0 / δ κ
Kρ m c p tanh[(1 + i ) y0 / δ κ ]
εs = . (2.1.34)
K s ρ s cs tanh[(1 + i )l / δ s ]

Real part
1.2 Imaginary part

1.0

0.8
tanh(1+i)y0/δk

0.6

0.4

0.2

0.0
0 1 2 3
-0.2 y0/δk

Figure 2.1.4: The real and imaginary parts of tanh (1 + i ) 


y0
 δ κ 

In Fig.2.1.4, the real and imaginary parts of tanh[(1 + i ) y0 / δ κ ] are plotted. Note
that at y0 = 2δ κ the function is almost unity.
16 Chapter 2

5 cm 5 cm 15 cm

0 x

Figure 2.1.5: Geometry of the example case.

As an illustration of the temperature oscillation for the T1 profile, an example is


plotted in Fig 2.1.6 and 2.1.7 based on Eq. (2.1.30). Here, the real part of the T1
variation along the y direction at any position in the stack is shown. In this example,
the same parameters as for the x component u1 in Fig.2.1.3 are adopted. The
geometrical schematic is shown in Fig. 2.1.5. The resonator tube is a quarter of one
wave length (25cm) and the stack length is one fifth of the resonator tube length
(5cm). The leading end of the stack is placed at 5 cm away from the pressure node.
The total energy flow E& 2 along the stack is zero. At a fixed x position, the real and
imaginary part of T1 increases as approaching the center of the fluid layer. At a
fixed y position, T1 decreases as approaching the anti-node of the pressure wave.
The relative position in the stack is evalued as x/stack length in Fig 2.1.6 and 2.1.7.
From Eq. (2.1.30), it is obvious that it consists of three groups of terms. For
convenience, name them as:
p1
the first term: T1 = ; (2.1.35)
ρ mc p
the second term:
1  σ cosh[(1 + i ) y / δ v ]  dp1 dTm
T1 = − × 1 −  ;
ρ mω  (σ − 1) cosh[(1 + i ) y0 / δ v ]  dx dx
2 (2.1.36)

and the third term:


 p
T1 = −  1 +
(dp1 dx )(dTm / dx ) 1 + ε s f v  × cosh[(1 + i ) y / δ κ ] .
 ρ m c p (σ − 1)ρ mω 2  f k  (1 + ε s )cosh[(1 + i ) y0 / δ κ ]
(2.1.37)
The first term comes from the adiabatic acoustic compressions and expansions. The
second and the third terms come from the oscillatory movement of the fluid along
the mean-temperature gradient in the fluid, with viscous effects included.
Basic theory of thermoacoustics 17

-2.000
0 -1.725

-1.450

-1.175

real part of T1 (K)


-0.9000

-0.6250

-0.3500

-0.07500

0.2000

1.0
0.8
-2
0.1 0.6
0.2
0.3 0.4
relat 0.4 0.5

y/y
0
ive p 0.6 0.2
ositio 0.7
n in 0.8
the s 0.9
1.0 0.0
tack

Figure 2.1.6: Real part of T1 at various y position and x position in the stack.

0.6

-0.2500
T (K)

-0.1312
0.4
imaginary part of 1

-0.01250

0.1063

0.2250
0.2 0.3438

0.4625

0.5813
0.0
0.7000

1.0
0.8
-0.2
0.1 0.6
0.2
0.3 0.4
0.4
y
0

0.5
y/

relat 0.6 0.2


i ve p 0.7
ositio 0.8
0.9
n in 1.0 0.0
th e s
tack

Figure 2.1.7: Imaginary part of T1 at various y position and x position in the stack.
18 Chapter 2

Wave equation
Next, the wave equation for p1 ( x) is derived. Starting with the continuity equation
∂ρ
+ ∇ ⋅ (ρV ) = 0 .
v
(2.1.38)
∂t
Substitution of the variables Eq. (2.1.1) to (2.1.6) into Eq. (2.1.38), and keeping the
first order terms yields

iωρ1 + (ρ mu1 ) + ρ m ∂v1 = 0 . (2.1.39)
∂x ∂y
Using Eq. (2.1.19), it can be written as:
1 dp1 µ ∂ 2u1
ρ mu1 = − + . (2.1.40)
iω dx iω ∂y 2
Substitution of Eq. (2.1.40) into Eq. (2.1.39) gives
d 2p1 ∂  µ∂ 2u1  ∂v
− ω 2ρ1 − +  2 
 + iωρ m 1 = 0 . (2.1.41)
dx 2
∂x  ∂y  ∂y
Assuming ideal-gas behavior, we can write:
ρ = p / RT (2.1.42)
where the specific gas constant is R = Runiv / m ,universal gas constant
Runiv = 8.3 J / mol ⋅ K , and m molecular weight.
Thus, we have
dρ = dp / RT − [( p / RT ) / T ]dT . (2.1.43)
Substitute the adiabatic speed of sound a = γRT [39], where γ is the ratio of
2

isobaric to isochoric specific heats. We find


(
ρ1 = − ρ m β T1 + γ a 2 p1 . ) (2.1.44)
Substitution of Eq. (2.1.44) in Eq. (2.1.41) yields
ω2 d 2p1 ∂  ∂ 2u1  ∂v
ω 2 ρ m βT1 − γ p1 − +  µ 2  + iωρ m 1 = 0 . (2.1.45)
a2 dx 2
∂x  ∂y  ∂y
Integrating Eq. (2.1.45) with respect to y from 0 to y0 yields a wave equation for
the first-order acoustic pressure amplitude p1 ( x)
y0 y0 ω2 y0 d 2p1 y0 ∂  ∂ 2u1 
∫ ω 2 ρ m βT1dy − ∫ γp1dy − ∫ dy + µ
∫ 0 ∂x  ∂y 2 dy

0 0 a2 0 dx 2
y0 ∂v1
+ ∫ iωρ m dy = 0 . (2.1.46)
0 ∂y
For the last term on the left hand side follows
y0 ∂v1
∫ iωρ m dy = iωρ m v1 y = y − iωρ m v1 y = 0 = 0 . (2.1.47)
0 ∂y 0
Basic theory of thermoacoustics 19

Because the boundary conditions: at y = y0 , due to the wall, v1 = 0 and at


y = 0 , v1 = 0 by symmetry.
Thus, Eq. (2.1.46) reduces to
y0 y0 ω2 y0 d 2p1 y0 ∂  ∂ 2u1 
∫ ω 2 ρ m βT1dy − ∫ γp1dy − ∫ dy + µ
∫0 ∂x  ∂y 2 dy = 0 .
 (2.1.48)
0 0 a2 0 dx 2
By substituting Eq. (2.1.30) for T1 , the first integration term is obtained
y0 ω 2β  f 
∫ ω 2 ρ m βT1dy = y0 1 − κ p1
0 cp  1+ εs 
dp1 dTm  σf v fκ + ε s f v 
− β y0 1 − + . (2.1.49)
dx dx  σ − 1 (σ − 1)(1 + ε s ) 
By substituting Eq. (2.1.20) for u1 , the last term in the left hand side of Eq.
(2.1.48) is obtained
y = y0
∂  ∂ 2u1 
y0 ∂ y0 ∂ 2u1 ∂  ∂u1 
 µ  dy µ
∫0 ∂x  ∂y 2  ∂x ∫0 ∂y 2
= dy = µ 
∂x  ∂y 
 y =0

d  dp 
= y0  1 ⋅ f v  . (2.1.50)
dx  dx 
Substitution of Eq. (2.1.49) and (2.1.50) into Eq. (2.1.48) yields
ω 2 βy0  f  dp dT  σ f v fκ + ε s f v 
1 − κ p1 − βy0 1 m 1 − + 
cp  1 + ε s  dx dx  σ − 1 (σ − 1)(1 + ε s ) 
ω2 d 2p1 d  dp 
− 2 γ y0p1 − y0 2
+ y0  1 ⋅ f v  = 0 . (2.1.51)
a dx dx  dx 
By using the following relations,
γR
cp = ; (2.1.52)
γ −1
1
β= ; (2.1.53)
T
a 2 = γ RT . (2.1.54)
Eq. (2.1.51) can be rewritten as
 (γ − 1) fκ

p1 + 2  (1 − fν ) 1 
a2 d dp
1 +
 1+ εs  ω dx  dx 
a 2 dT dp  σf fκ + ε s fν 
+ β 2 m 1 1 − v + = 0.
ω dx dx  σ − 1 (σ − 1)(1 + ε s ) 
(2.1.55)
20 Chapter 2

Using the state equation (2.1.42), the second term of Eq. (2.1.55) on the left hand
side can be written as
a2 d  dp1  ρ m a 2 d 1 − f v dp1  a2
 (1 − f )  =   − β (1 − f v ) dTm dp1 .
ω dx 
2 v
dx  ω dx  ρ m dx 
2
ω 2
dx dx
(2.1.56)
Substituting Eq. (2.1.56) into (2.1.55), the thermoacoustic wave equation is
obtained
 (γ − 1) fκ  ρ m a 2 d  1 − fν dp1  a2 fκ − fν dTm dp1
1 + p1 + 2   − β 2 =0.
 1+ εs  ω dx  ρ m dx  ω (1 − σ )(1 + ε s ) dx dx
(2.1.57)
This equation describes a sound field modified by the interaction between fluid and
solid plates. The coefficients, related to the acoustic pressure and its gradient, are
complicated functions having a dependence on the temperature profile. If the
temperature profile is known, the wave equation (2.1.57) can be solved. In chapter
4, this wave equation is reduced by some assumptions to describe the sound field
inside the different components of a traveling-wave system.

The time-averaged total energy flow

We will proceed to derive an expression for the time-averaged energy flow. In


steady state, and assuming that the system is ideally isolated from the surroundings,
it can be deduced that the time-averaged energy flow must be independent of x.
Start with conservation of energy [38]:
∂ 1 v2  v  v 1 v 2  v v v
 ρ V + ρe  = −∇ ⋅  ρV  V + h  − K∇T − V ⋅ ∑ , (2.1.58)
∂t  2   2  
where e and h are internal energy and enthalpy per unit mass, respectively, and
∑ is the viscous stress tensor, with components:
 ∂v ∂v 2 ∂v  ∂v
∑ ij = µ  i + j − δ ij k  + ξδ ij k . (2.1.59)
 ∂x 
 j ∂xi 3 ∂xk  ∂xk
The first term on the left hand side in Eq. (2.1.58) is kinetic energy density of unit
control volume. The second term on the left hand side is the internal energy density.
The first term on the right hand side is from the enthalpy flow and kinetic energy.
The second term is from thermal conduction. The last term on the right hand side
results from the viscosity. The lowest-order variation in the energy is of second
v v2
order. All terms of higher order (e.g. V V are of third order) are neglected.
Basic theory of thermoacoustics 21

Integrating the remaining terms in Eq. (2.1.58) with respect to y from y=0 to
y′ = 0 and time averaging, yields
d  y0
dx 
y0
 ∫ 0 ρuhdy − ∫ 0 K
∂T
∂x
l ∂T y0 v v 
dy − ∫ K s s dy′ − ∫ V ⋅ ∑ x dy  = 0 .
∂x
( ) (2.1.60)

0 0

The over bar denotes time averaging. The quantity within the square brackets is the
time-averaged energy flow per unit of perimeter along x, defining this quantity as
E& ∏ , where Π is the perimeter of the stack plates:
E& y0 y0 ∂T l ∂T y0 r v
= ∫ ρuhdy − ∫ K dy − ∫ K s s dy′ − ∫ (V ⋅ ∑) x dy , (2.1.61)
∏ 0 0 ∂x 0 ∂x 0

where E& is the total energy flow through the stack and ∏ is the total perimeter of
the stack plates. Now h can be expanded in the same way as in the Eqs. (2.1.1) to
(2.1.6):
h = hm ( x) + Re[h1 ( x, y)eiω t ] (2.1.62)
Substitution of the equations from Eq.(2.1.1) to (2.1.6) and (2.1.62) into (2.1.61)
and expanding E& ∏ to second order in the acoustic amplitude, (the variation
terms of third order and higher are again neglected) the first term in Eq. (2.1.61)
becomes


y0

0
ρuhdy = ∫
0
y0
(ρ h Re[u e
m m 1
iω t
] + ρ m hm Re[u 2e 2iω t ] + hm Re[ρ1eiω t ] Re[u1eiω t ]

+ ρ m Re[u1eiω t ] Re[h e ])dy .


1
iω t
(2.1.63)
It is easy to see that the value of time averaging of the first term is also zero, i.e.
Re[u1eiω t ] = 0 .
The integrals of the third and fourth terms in Eq. (2.1.63) sum to zero because the
second-order time-averaged mass flow is zero:
y0
∫ ( ρ m Re[u 2 e 2 iω t ] + Re[ρ1eiω t ] Re[u1eiω t ])dy = 0 . (2.1.64)
0

Hence, the Eq. (2.1.63) reduces to


y0 y0
∫ ρuhdy = ∫ ρ m Re[u1eiω t ] Re[h1eiω t ]dy . (2.1.65)
0 0

Using equation
dh = Tds + (1 / ρ )dp (2.1.66)
and Eq. (2.1.26) yields
dh = c p dT + (1 / ρ )(1 − Tβ )dp . (2.1.67)
By using Eq. (2.1.67), the Eq. (1.1.65) becomes

∫0
y0 y0
[
ρuhdy = ∫ ρmc p Re[T1eiω t ] Re[u1eiω t ] dy .
0
] (2.1.68)
22 Chapter 2

For the second and third integrals of Eq. (2.1.61), only the zero order terms are
significant to be counted. The terms to second order or higher can be neglected.
Therefore, the two integrals are
∂T ∂T
dy − ∫ K s s dy′ ≅ −( y0 K + lK s ) m .
y0 l dT
−∫ K (2.1.69)
0 ∂x 0 ∂x dx
Using arguments similar to those leading to Eq. (2.1.19), we find that the largest
terms in the last integral in Eq. (2.1.61) are of order y0 µu1 / D , whereas ρuh has
2

the order of p1u1 ≅ ρ m au12 . Hence,

∫ (V ⋅ ∑ ) dy ∫
v
y0 v y0 ν 1 δ v2
x ρuh dy ~ = << 1 . (2.1.70)
0 0 Da 2 D2
(v )
v
So, the viscous term V ⋅ ∑ x is negligible. Therefore, Eq. (2.1.61) becomes
E& 2
∏ 0
y0
[ ]
= ∫ ρmc p Re[T1eiω t ] Re[u1eiω t ] dy − ( y0 K + lK s ) m .
dT
dx
(2.1.71)

To remind us that it is an energy flow valid to second order in the acoustic


quantities, here, the subscript “2” is added to E& .
Substitution of Eq. (2.1.30) for T1 and Eq. (2.1.20) for u1 , and integration yields
(see Appendix D)

E2 =
& ∏ y0
Im 
 dp
~  ~
p1 1 − fν −
(
fκ − fν
~
) 

(1 + ε s )(1 + σ ) 
1
2ωρ m  dx 

+
∏ y0 c p dTm dp1 dp ~
1
~
× Im  fν +
( )
fκ − fν (1 + ε s fν / fκ ) 
~

2ω 3 ρ m (1 − σ ) dx dx dx  (1 + ε s )(1 + σ ) 
dT
− ∏( y 0 K + lK s ) m , (2.1.72)
dx
where the tilde denotes complex conjugation. This equation describes up till second
order in the acoustic quantities energy flow in a thermoacoustic stack.

2.2 Acoustic energy

Now, we will develop an expression for the time-averaged acoustic power W& used
or produced in a segment of length ∆x in the stack. It is clear that the time-
averaged products of first-order terms in pressure and acoustic particle velocity are
the largest non-zero time-averaged power component. In acoustics, this is called
the acoustic intensity, which describes the time-averaged “rate per unit area at
which work is done by one element of fluid on an adjacent element” [39]. This
Basic theory of thermoacoustics 23

acoustic power is a second-order quantity. Therefore, the subscript “2” is added to


the power symbol W& . According to the geometry in Fig. 2.1.1, no net acoustic
power flows in the y direction. Thus, the difference in time-averaged acoustic
intensity between the two positions along the stack must be the acoustic power per
unit area generated or absorbed between those two positions. It is shown in Fig.
2.2.1.

Sound wave
x

Left Right

Figure 2.2.1: Two faces in a multi-plate thermoacoustic system used for


computation of the acoustic power flow.

Consider a cross section of the stack at position x and one at position x + dx .


Then the difference in acoustic power flow through these cross sections is given by
dW&2 = W&2 ( x + dx) − W&2 ( x)

[( ) ] [( )]
y0 y0

= ∏ ∫ Re[p1eiω t ] Re[u1eiω t ] ( x +dx ) dy − ∏ ∫ Re[p1eiω t ] Re[u1eiω t ] ( x ) dy .


0 0
(2.2.1)
When dx is infinite small, we can write

∫ (Re[p e )
y0
d iω t
dW&2 = ∏ dx 1 ] Re[u1eiω t ] dy ,
dx 0
Or
d  
y0
dW&2
=∏ Re[p1e ] ∫ Re[u1eiω t ]dy  .
iω t
(2.2.2)
dx dx  
 0 
The velocity in the x-direction, averaged over the cross section is defined as:
y
1 0
y0 ∫0
u1 = u1dy (2.2.3)

Eq. (2.2.2) can be rewritten as:


dW&2
dx
= ∏ y0
d
dx
(
Re[p1eiω t ] Re[ u1 eiω t ] . ) (2.2.4)

By using Eq. (D.2), Eq. (2.2.4) can be expressed as


24 Chapter 2

~
dW&2 1 ~ ] = 1 ∏ y Re p d u1 + u
= ∏ y0 Re[p1 u
d ~ dp1  .
1 0  1 1  (2.2.5)
dx 2 dx 2  dx dx 
According to Eq. (2.1.20), the average velocity can be obtained as:
u1 =
i

dp1
(1 − fν ) . (2.2.6)
ωρ m dx
The conjugate average velocity is
( )
~
~ = − i ⋅ dp1 1 − ~
u fν . (2.2.7)
1
ωρ m dx
From Eq. (2.2.6), it can be obtained
dp1 − iωρ m u1
= . (2.2.8)
dx 1 − fν
~ / dx can be obtained as:
From Eq. (2.2.7), d u1
~
d u1
= 
~ ~
(
− i d  1 − fν dp1

.
) (2.2.9)
dx ω dx  ρ m dx 
According to Eq. (2.1.57):
d  (1 − fν ) dp1  ω 2  a2 fκ − fν dTm dp1  (γ − 1) fκ  
  = 2 
β 2 − 1 + p1  .
dx  ρ m dx  ρ m a  ω (1 − σ )(1 + ε s ) dx dx  1+ εs  
(2.2.10)
Substitution of Eq. (2.2.10) in (2.2.9) yields:
~ ~ ~ ~ ~
d u − iβ fκ − fν dTm dp iω  (γ − 1) fκ ~
1
= 1
+ 1 + p1 . (2.2.11)
dx ωρ m (1 − σ )(1 + ε~s ) dx dx ρ m a 2  1 + ε~s 

Substituting Eq. (2.2.8) and Eq. (2.2.11) into Eq.(2.2.5), gives:

(γ − 1) p1 2  fκ 
2
dW&2 1 ρ u
~ 2 Im( fν ) + ρ a 2 Im 1 + ε 
= ∏ y0ω  m 1
 
dx 2 
 1 − fν
m  s 

~ ~
1 β dTm  ( fκ − fν ) ~ .
+ ∏ y0
(1 − σ ) dx  (1 + ε~s )(1 − fν ) 1 
Re  ~ p 1 u (2.2.12)
2
This is the acoustic power absorbed (or produced in the prime mover mode) in the
stack per unit length.
The first two terms in Eq. (2.2.12) are the viscous and thermal relaxation
dissipation terms, respectively. These two terms have a dissipative effect in
thermoacoustics and they will be present whenever a wave interacts with a solid
surface. The third term can be either source or sink for acoustic power. It depends
on the sign of the temperature gradient along the stack.
Basic theory of thermoacoustics 25

This chapter is dedicated to the review of Rott’s theory. The time-averaged total
energy flow Eq.(2.1.72) is an important equation to be used often in the latter
chapters.
Chapter 3

Standing-wave systems

3.1 Introduction

Thermoacoustics employs the thermodynamic interaction between a solid wall and


sound wave to realize heat transport (in heat pump mode, taking heat from cold end
to the warm end) or energy conversion (in prime mover mode, changing the heat
into mechanical energy—acoustic energy). In chapter 2, it is shown that these
thermoacoustic effects occur in a thin gas layer, which is called the fluid’s thermal
penetration depth. In a standing-wave system, these thermoacoustic effects are used
by inserting a stack of parallel plates made from poorly conductive materials into a
resonator tube. The plates are normally spaced by a distance about twice the
working gas’ thermal penetration depth. Thus, most of the gas between the stack
plates takes part in the heat transport. This is the basic idea behind the construction
of a standing-wave thermoacoustic refrigerator. Normally, a standing-wave
thermoacoustic refrigerator consists of an acoustic driver to generate a sound wave,
a resonator tube filled with a working gas and a stack sandwiched between a cold-
end heat exchanger and a hot-end heat exchanger. In practice, the resonator tubes
are made in a quarter wave-length or in a half wave-length. These two types of
standing-wave systems are illustrated schematically in the figures 3.1.1 a and b,
respectively. In a quarter wave-length resonator, there is a large-volume gas
reservoir at one end to create the right acoustic boundary condition at the end of the
resonator. In a half wave-length resonator, one end is closed by the acoustic driver
and the other one is just a closed end.
The foundation of linear thermoacoustic theory by Rott [1] unveiled a very prolific
era of thermoacoustics. This novel technology attracted much attention and interest
from researchers. Among them, the Condensed Matter and Thermal Physics Group
of Los Alamos National Laboratory (LANL) are the distinguished pioneers. In the
Standing-wave systems 27

early 1980s, they dedicated their efforts to the exploration of thermoacoustic


concepts to create devices that would produce useful refrigeration or useful work
[40-42].

Sound wave Stack Resonator tube

Ambient Cold
Loudspeaker HX HX
Gas reservoir

Figure 3.1.1: (a) A simple illustration of a quarter wave-length standing-wave


thermoacoustic refrigerator.

Cold HX Ambient HX
Sound wave
Loudspeaker Resonator tube Stack

Figure 3.1.1: (b) A simple illustration of a half wave-length standing-wave


thermoacoustic refrigerator.

Tom Hofler’s PhD work was part of these efforts, who built the first efficient
thermoacoustic refrigerator [43]. Hofler’s standing-wave refrigerator was a quarter
wave-length one, having a similar configuration as shown in Fig. 3.1.1a.
A standing-wave type refrigerator was launched on the space shuttle Discovery in
1992 [44]. It was built in Naval Postgraduate School (NPS), designed to produce
up to 80 K temperature difference over the stack, and to pump up to 4 W of heat.
Another refrigerator built in NPS was the Shipboard Electronics Thermoacoustic
Cooler to cool radar electronics on board of the warship USS Deyo in 1995 [45]. It
was able to provide 400 W of cooling power for a small temperature span as
designed.
At Pennsylvania State University, a large refrigerator called TRITON was built for
cooling of Navy ships, which was designed to generate a cooling power of 10 kW
[45].
At LANL, a heat-driven thermoacoustic refrigerator “beer cooler” was built, which
replaced the only moving part, the diaphragm of the loudspeaker, by
thermoacoustic prime mover. It has no moving parts at all. Two of similar devices
28 Chapter 3

were also built at NPS: one is called a Thermoacoustic Driven Thermoacoustic


Refrigerator (TADTAR), having a cooling power of about 90 watt for a
temperature span of 25°C; The other TADTAR is solar driven, having a cooling
capacity of 2.5 watt for a temperature span of 17.7°C [47].
After this general introduction of standing-wave system, the following subsections
are dedicated to: a simplified physical description of standing-wave system
thermoacoustic cycle; modeling of standing-wave systems; and a standing-wave
type experiment apparatus, which is similar to the so-called “TAC”
(thermoacoustic couple), was built to validate the model developed in section of
modeling. At the last part of this chapter, the measurement results are analyzed and
compared with computation based on proposed modeling.
A “TAC” was introduced for better understanding of heat transport inside a
standing-wave system for its simply structure. With more work were carried out by
researchers on similar apparatus at high amplitudes, large discrepancies between
measurement and computation by using the linear thermoacoustic theory were
found and reported. These deviations from linear thermoacoustic theory attracted
much attention and a succession of works was conducted to explore the attributions.
In this chapter, a new approach to evaluate the total energy flow along the stack
and then to further compute other parameters about the system was proposed and
tested by measurements.
Standing-wave systems 29

3.2 Physical description of standing-wave systems

The basic principles of the thermoacoustic effect in a standing-wave type


refrigerator will be given in this section. In chapter 2, the basic theory of
thermoacoustics is developed from an Eulerian point of view of fluid mechanics. In
this section, to make it easier to understand, a Lagrangian point of view is adopted
[53]. A given typical parcel of the fluid will be traced as it moves. The sound wave
that is sinusoidal in practice is simplified as a square wave. By these
simplifications, the cycle is simplified to two reversible adiabatic steps and two
irreversible constant-pressure steps, which is called the Brayton cycle. The
animation of this thermal process is clearly illustrated at the Los Alamos National
Laboratory web site [54].
This thermoacoustic cycle is schematically presented in Fig. 3.2.1. A given gas
parcel oscillating along the stack plate at the distance of a fluid’s thermal
penetration depth is traced. A longitudinal temperature gradient ∇Tm is assumed to
exist along the stack. Additionally, it is supposed that the pressure antinode is to
the right end of the plate and a node to the left. For simplicity, an inviscid ideal gas
is assumed to be the working fluid. The four steps of the thermoacoustic cycle are
shown separately in Fig. 3.2.1 (a). A starting point is set as the position where the
traced gas parcel is at the left most position of its stroke in this oscillation. The
temperature, pressure, and volume of the parcel are Tm − x1∇Tm , p m − p1 , and
V respectively. In step 1, the gas parcel moves a distance 2x1 , where it reaches its
right most position. The given gas parcel is compressed by the periodic acoustic
wave oscillation and the inside temperature is increased by an amount of 2T1 . The
pressure change 2 p1 and the temperature change 2T1 are related by the
thermodynamic relationship
 ∂T  1  ∂ρ  1  ∂ρ   ∂T  Tm β
T1 =   p1 = − 2   p1 = − 2     p1 = p1
 ∂p  s ρ m  ∂s  p ρm  ∂T  p  ∂s  p ρ mc p
(3.2.1)
where s is entropy per unit mass, β = −(∂ρ / ∂T ) p / ρ m is the ordinary thermal
expansion coefficient, and c p is the constant-pressure heat capacity per unit mass.
After step 1, the temperature, pressure and volume become
Tm − x1∇Tm + 2T1 , p m + p1 and V − V1 respectively. At the same time, the gas
parcel is located at the right most position, where the local temperature of the stack
plate is Tm + x1∇Tm . At this moment, the temperature difference between the gas
parcel and the stack plate is
δT = 2T1 − 2 x1∇Tm . (3.2.2)
30 Chapter 3

a) 1. Adiabatic compression 2. Heat transfer

Tm − x1∇Tm Tm + x1∇Tm
Plate
δQ1
2x1
δκ δW1 δW2
Gas
Parcel pm − p1 pm + p1 pm + p1
Tm − x1∇Tm Tm − x1∇Tm + 2T1 Tm + x1∇Tm
V V − V1 V − V1 − δV

3. Adiabatic expansion 4. Heat transfer

δQ2
2x1
δW3 δW4

pm − p1 pm + p1 pm − p1
Tm + x1∇Tm − 2T1 Tm + x1∇Tm Tm − x1∇Tm
V − δV V − V1 − δV V

b) TC LS TH

Pressure Plate Pressure


node end QC QH antinode
δQ end
δκ x1

Figure 3.2.1: (a) Typical gas parcel in a thermoacoustic refrigerator mode,


experiencing a four-step cycle with two adiabats (step 1 and 3) and two constant-
pressure heat transfer (step 2 and 4). (b) An amount of heat is shuttled along the
stack plate from one gas parcel to the next, as a result, heat Q is transported from
the left end of the plate to the right end, using work W [53].

In a heat pump mode, there must be a positive δT . Therefore, in step 2, assuming


that the gas cools down to the local temperature of the stack the heat that is
transferred from the gas parcel to the stack plate is
δQ = mc p δT , (3.2.3)
where m is the mass of the gas parcel.
A schematic pV-diagram of the cycle is shown in Fig.3.2.2. Here, the adiabatic
steps 1 and 3 are linearized which is allowed because of the small δV . The work
used to drive this cycle is equal to the area ABCD, is given by
Standing-wave systems 31

δW = ∫ pdV (3.2.4)
ABCD

The used work can be seen from Fig. 3.2.2. Hence, the work used in this 4-step
cycle is given by
δW ≈ −2 p1δV . (3.2.5)

Pressure

C 2 B
pm + p1

1
3

pm − p1 A
D 4

V − V1 − δV V − V1 V − δV V Volume

Figure 3.2.2: Schematic pV-diagram of the thermoacoustic cycle of Fig. 3.2.1. The
four steps of the thermoacoustic cycle are illustrated: adiabatic compression 1,
isobaric heat transfer 2, adiabatic expansion 3 and isobaric heat transfer 4. The
area ABCD is the work used in the cycle [53].

After releasing the heat from the gas parcel to the stack plate, the gas parcel has the
same temperature as the local stack plate and in this isobaric heat transfer process
its volume is reduced. In step 3, the gas parcel moves back to the initial most left
position via an adiabatic expansion. The pressure drops to p m − p1 by the
periodical sound movement. The accompanying temperature drop is 2T1 by this
adiabatic expansion. The volume expands by V1 to V − δV . After this expansion,
the gas parcel becomes colder than the local stack plate. In step 4, the gas parcel
absorbs heat from the local stack plate to eliminate the temperature difference
− δT by an isobaric expansion. In the two isobaric processes, step 2 and 4, the
temperature difference between the gas parcel and the local temperature of the
stack plate is crucial to the direction of the heat flow. If the δT is positive, then, in
step 2, the gas parcel has a higher temperature than that of the stack plate, and in
step 4, the gas parcel has a lower temperature. By this means, the gas parcel
absorbs heat in the left most position and release heat at the extreme right position,
realizing the heat transportation from left to right. If the δT is negative, the whole
story reverses. The gas parcel absorbs heat from the stack plate when it arrives at
the right most position and gives off heat as it is at the extreme left position. This
32 Chapter 3

process heats up the gas parcel while compressed, and cools it down while
expanded. This makes the sound amplified and is called “prime mover” mode. If
the δT is zero, then no heat is exchanged between the gas parcel and the stack plate.
Therefore, there exists a critical temperature gradient, which distinguishes the “heat
pump” mode from the “prime mover” one. The critical temperature gradient is
given by δT = 0 , when the temperature change 2 x1∇Tm by the gas parcel
displacement from the sound wave transportation is just matched by the adiabatic
temperature change 2T1 by the sound wave thermodynamic process. Eq. (3.2.2)
yields
∇Tm critical = ∇Tcrit = T1 / x1 (3.2.6)
Using Eq. (3.2.1) and x1 = u A / ω , where u A is the gas particle velocity amplitude
at the specific position along the stack and ω the angular frequency, it is obtained
Tm βωp A
∇Tcrit = . (3.2.7)
ρmc pu A
Usually, the displacement of a given gas parcel is small with respect to the length
of the plate. Thus there will be an entire train of adjacent gas parcels, each confined
in its cycle motion. For the heat pump mode, during the first half cycle, the
individual gas parcels will each move a distance of 2x1 towards the pressure
antinode and deposit heat δQ locally on the stack plate. During the second half
cycle, each parcel moves back to its initial position and picks up heat δQ from the
plate. But this heat was deposited there a half cycle earlier by an adjacent gas
parcel. Eventually, the heat δQ is passed along the plate from one gas parcel to the
next one in the direction of the pressure antinode. By means of these gas parcels in
series, the heat δQ is shuttled along the plate towards the pressure antinode end.
The ratio of the maximum acoustic pressure amplitude in the system and the mean
pressure is called the “drive ratio”.

3.3 Modeling standing-wave systems

In thermoacoustics, thermodynamics is strongly coupled with acoustics. To


compute the thermoacoustic effects, it is necessary to find information of the
acoustic field, which will be modified by the thermodynamic effects induced by the
presence of a “stack”. The interaction between thermodynamics and acoustics is
very complicated. In order to simplify the analysis, we, therefore, decouple the
acoustics from the thermodynamics. For a standing-wave machine, the decoupling
is realized by assuming that a standing wave acoustic field exists in the resonator,
Standing-wave systems 33

without being influenced by the presence of the stack. Normally, the sound field
consists of two components: standing-wave and traveling-wave component. This
can be seen in appendix E. In most practical cases, the traveling-wave component
is much smaller than the standing-wave component. Therefore, to make a further
simplification, we assume that the traveling-wave component is negligible.
In a standing-wave system, the total energy flow along the stack E& 2 is an important
parameter, which can be deduced by the analysis of the energy balance of the
system. In the sections 3.3.1 and 3.3.2, the total energy flow along the stack E& 2 is
assumed to be known. Both the total energy flow along the stack E& 2 and the
viscosity of the working gas have much influence on the performance of the whole
system. Therefore, the computations in sections 3.3.1 and 3.3.2 are arranged under
categories by using different combinations of viscosity and total energy flow E& 2
(zero viscosity and zero total energy flow E& 2 =0; zero viscosity and non-zero total
energy flow E& 2 ≠0; non-zero viscosity and zero total energy flow E& 2 =0; non-zero
viscosity and non-zero total energy flow E& 2 ≠0 ). By doing so, the performance ∆T
and temperature gradient dTm/dx as a function of stack position are investigated in
a manner of step by step from simple to complicated.

3.3.1 Zero viscosity

First consider zero viscosity and the so-called “short-stack” approximation:


assume that the stack plates do not disturb the sound field.
The assumption of “zero viscosity” makes the axial flow velocity amplitude
u1 independent on y. So, u1 is uniform between two adjacent stack plates.
Furthermore, we use the “boundary-layer” approximation which means y0 >> 2δ κ
and l >> 2δ s , so that the hyperbolic tangents ( tanh[(1 + i ) y0 / δ κ ] ) in Eqs. (2.1.31),
(2.1.32) and (2.1.33) can be set to unity.
The acoustic field inside of the resonator is
p1 = p A sin( x / D ) ≡ p1s , (3.3.1)
u1 = i (1 + l / y0 )( p A / ρ m a )cos( x / D ) ≡ iu1s . (3.3.2)
Here, p A is the pressure amplitude in the pressure anti-node. The term (1 + l / y0 )
accounts for the continuity of the volume flow at the ends of the stack.
With all these assumptions applied to Eq. (2.1.72), the formula to compute the
energy flow is as follows:
p1su1s (Γ − 1) − ∏( y0 K + lK s ) m ,
1 1 dT
E& 2 = − ∏ δ κ (3.3.3)
4 1+ εs dx
34 Chapter 3

where
ρ m c pδ κ
εs = . (3.3.4)
ρ s csδ s
In Eq. (3.3.3), the ratio between the actual temperature gradient and the critical
gradient is represented by
∇Tm
Γ= .
∇Tcrit (3.3.5)

If the conductive loss is neglected (the second term on the right hand side in Eq.
(3.3.3)) and E& 2 = 0 , then all available acoustic power is used to establish a
temperature gradient. This maximum attainable temperature gradient is the critical
temperature gradient given by Eq. (3.2.7), determined by the cyclic compression
and expansion of the gas.
Substituting Eq. (3.3.1), (3.3.2) into (3.2.7) yields
ωa x
∇Tcrit = tan .
c p (1 + l / y 0 )
(3.3.6)
D
In Eq. (3.3.6), the critical temperature gradient depends on the operating frequency
and the geometrical properties of the system. Assuming that the working gas is
helium, the working frequency is 1000 Hz at ambient temperature, and that the
thickness of the stack plate is much smaller than the working gas layer i.e.
l / y 0 ≈ 0 . The critical temperature gradient plot in that case is shown in Fig.3.3.1.
7

6
Log(critical temperature gradient)

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
x/radian wavelength

Figure 3.3.1: Critical temperature gradient given by Eq. (3.3.6).


Standing-wave systems 35

Remark that in the position, x/radian wavelength=π/2, where a pressure anti-node


is occurring the critical temperature gradient becomes infinity due to the tan
function. At this position, steep gradient may occur. In the pressure node, at x=0,
the temperature gradient is zero.

Zero total energy flow


Now, we focus on the situation when the total energy flow given by Eq. (3.3.3) is
zero, which means that all the thermal power produced by the acoustic wave is
consumed via conduction by the working gas itself and by the stack plates. So a
temperature gradient is established without cooling power output. Also, we are
assuming the working fluid is ideal gas. The length of the stack in all computations
in this chapter is not so short that the temperature gradient can be considered as
constant. Therefore, the temperature gradients depend on positions.
Then from Eq. (3.3.3), the temperature gradient in that case can be expressed as:
dTm ωδ κ p1s u1s
= .
dx 4ω (1 + ε s )( y 0 K + lK s ) + δ κ ρ m c p u1s ( ) 2 (3.3.7)

Substitution of the acoustic pressure Eq. (3.3.1) and the particle velocity Eq. (3.3.2)
into Eq. (3.3.7) leads to
dTm sin (2 x / D )
= ,
B + D ⋅ cos(2 x / D )
(3.3.8)
dx
Where
 8ω (1 + ε s )( y 0 K + lK s )ρ m a 2  c p (1 + l / y 0 )
B= + 1⋅ ,
δ κ c p (1 + l / y 0 )2 p A2
(3.3.9)
  aω
and
c p (1 + l / y0 )
D= . (3.3.10)

The temperature difference between the starting point of the stack and the other
end of the stack is represented as:
x stack + l stack
x stack + l stack dTm ( x) − a2   2 x 
∆Tm ( x) = ∫ dx = ln  B + D⋅cos  .
x stack dx 2c p (1 + l / y0 )   D  x stack

(3.3.11)
In Eq. (3.3.11) xstack is the starting point of the stack and lstack is the length of the
stack. Now considered as an example case, a quarter-wave-length standing wave
machine, is shown in Fig. 3.3.2. The working frequency is assumed to be 1000 Hz.
The working gas is helium. The stack length is taken as 1/20 wave length, which is
short enough not to significantly disturb the sound field as assumed before (“short-
36 Chapter 3

stack” approximation). The plate thickness 2l is fixed to twice the thermal


penetration depth of the solid material. All the parameters of this example model
are listed in table 3.3.I.
Cold HX Hot HX
Cold HX

λ/4
x
0

Figure 3.3.2: Geometry of the standing-wave example case.

Working gas
Helium Properties

Operating frequency 1000 Hz

Mean pressure 1 bar


1 kPa (1% of mean pressure) or
Acoustic pressure
amplitude 10 kPa (10% of mean pressure)

Ambient temperature 300K

Density 0.16 kg/ m 3


Thermal conductivity 0.156 w/(m*K)

Cp 5193 J/(kg*K)

speed of sound 1019 m/s


Stack plate
Properties
PVC
Thermal conductivity 0.2 W/(m*K)
Density 1300 kg/ m 3

Cp 1500 J/(kg*K)
Table 3.3.I (a) Properties of working fluid and stack of the standing-wave example
case
Standing-wave systems 37

Geometry parameters data

Thermal penetration depth (gas) 2.44e-4 m


Thermal penetration depth (solid) 5.71e-6 m

Stack length (1/20 wave length) 0.051 m

Tube length ( ¼ wave length) 0.255 m

Diameter of tube (1/20 wave length) 0.051 m

Plate thickness 11.5 µm


Table 3.3.I (b) Geometrical parameters of the standing-wave example case

x Stack position
xstack

Figure 3.3.3: Illustration of the stack position.

The heat exchangers are ideal. The left end of the resonator is where the pressure
node locates. The right end of the resonator is closed by a solid wall or a speaker,
where the pressure anti-node locates.
The influence of the stack-plate spacing on the performance of the system is
investigated, as well as that of the stack position in the resonator tube. Here the
stack position is referred to the end of the stack nearest to the open end of the tube,
see Fig. 3.3.3. The half distance between the stack plates y 0 is varied and set to the
values of 2δ κ 5δ κ 10δ κ 20δ κ and 50δ κ . Figs. 3.3.4 and 3.3.5 show the acoustic
pressure amplitude, and the acoustic particle velocity amplitude at different
positions inside the resonator tube for the case of a drive ratio of 10%. Obviously,
the pressure amplitude is maximum at the right end of the resonator (x=0.255m),
whereas the velocity amplitude is maximum at the left end of the tube (x=0).
38 Chapter 3

10000

Acoustic pressure amplitude (Pa)


8000

6000

4000 pm=1 bar, pA=10 kPa (10%)

2000

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.4: Acoustic pressure amplitude distribution inside the resonator,


drive ratio=10%.

70
Acoustic particle velocity amplitude (m/s)

60
pm=1 bar,pA=10 kPa (10%)
50

40

30

20

10

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.5: Acoustic particle velocity amplitude distribution inside the resonator,
drive ratio=10%
Standing-wave systems 39

16000

critical temperature gradient


14000
y0/δκ=2.0
y0/δκ=5.0
12000
y0/δκ=10.0
y0/δκ=20.0
dTm/dx (K/m)
10000
y0/δκ=50.0
8000

6000

4000

2000

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.6: Calculated temperature gradient on the stack at various positions


and for different stack plate spacing, drive ratio=10%.(Symbols are for indication
of different ratios of y0/δκ.)

Figure 3.3.6 shows that the temperature gradient decreases with increasing
thickness of the gas layer. This is due to the increasing thermal conduction by the
gas and plates while the amount of effective working gas is less. This effective
amount of gas basically is in the layer of one thermal penetration depth away from
the plate. This amount becomes less when the stack plate spacing increases because
the number of plates in a given cross section reduces. For any gas-layer thickness
in Fig. 3.3.6, the temperature gradient is always below the critical temperature
gradient. This is because in the critical gradient thermal conduction is not taken
into account. As the stack plate spacing decreases, the maximum local temperature
gradient moves towards the right end of the resonator tube where the pressure anti-
node lies. If the gas layer is thinner, the ratio of conductive heat loss to the power
generated by the gas becomes smaller. So, the situation will be closer to that of the
critical temperature gradient. In Fig.3.3.6, the curve of the thinnest gas layer
approaches the critical temperature curve closest. The critical temperature gradient
reaches infinity at the closed end of the resonator tube. This can be explained by
Fig. 3.3.5: the acoustic particle velocity is zero at the closed end of the resonator
tube, whereas the acoustic pressure amplitude is maximum (Fig. 3.3.4). Therefore,
40 Chapter 3

the temperature gradient goes to infinity at this position. In the case of conductive
flow Γ < 1 and dTm / dx becomes zero at the closed end since u1 = 0 (Eq. 3.3.3).
Since the temperature gradient in the stack material is equal to that in the gas, the
conductive heat flow can be expressed as
dTm
Π ( y0 K + lK s ) ⋅ . (3.3.12)
dx
Per unit of cross-sectional area the conductive heat flow equals to
y0 K + lK s dTm
. (3.3.13)
y0 + l dx
The temperature difference that is established between the two ends of the stack
can be determined by integrating the gradient as depicted in Fig. 3.3.6 along the
length of the stack. The resulting temperature difference for the example case of
table 3.3.I is shown in Fig. 3.3.7.

400
Temperature difference over the stack (K)

350
y0/δκ=2.0
300 y0/δκ=5.0
y0/δκ=10.0
250 y0/δκ=20.0
y0/δκ=50.0
200

150

100

50

0
0.00 0.05 0.10 0.15 0.20
Stack position (m)

Figure 3.3.7: Calculated temperature difference over the stack for different stack
positions and different stack plates spacing, drive ratio=10%.(Symbols are for
indication of different ratios of y0/δκ.)

The temperature span over the stack reaches its maximum value at some position
near the closed end of the resonator tube. As the gas layer becomes thinner, the
Standing-wave systems 41

stack starting position for maximum temperature difference over the stack moves
towards the closed end of the resonator tube. The thinner the gas layer, the larger
the temperature span. Although the temperature span is large at positions near the
closed end, the properties of the gas and solid are assumed to be independent on
temperature. So, all the physical parameters remain constant.
These evaluations were also performed for a drive ratio of 1%. In that case, the
temperature gradient developed in the stack obviously is much smaller as shown in
Fig.3.3.8. In addition the peak becomes wider. This is due to the fact that as p A
increases the parameter B in Eq. (3.3.8) becomes much smaller and the
denominator in the gradient is largely determined by the cosine term. For the same
reason, the temperature difference between the ends is smaller at smaller drive ratio,
see Fig. 3.3.9.

critical temperature gradient


y0/δκ=2.0
1600 y0/δκ=5.0
y0/δκ=10.0
1400
y0/δκ=20.0
1200
y0/δκ=50.0
dTm/dx (K/m)

1000

800

600

400

200

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.8: Calculated temperature gradient in the stack at various positions and
for different stack plate spacing, drive ratio=1%.(Symbols are for indication of
different ratios of y0/δκ.)
42 Chapter 3

60
y0/δk=2.0
y0/δk=5.0

Temperature difference over the stack (K)


50 y0/δk=10.0
y0/δk=20.0
y0/δk=50.0
40

30

20

10

0
0.00 0.05 0.10 0.15 0.20
Stack position (m)

Figure 3.3.9: Calculated temperature difference over the stack for different stack
positions and different stack plates spacing, drive ratio=1%.

Non-Zero total energy flow


In this section, the more general situation with non-zero total energy flow is
investigated. Substitution of Eq. (3.3.1) and (3.3.2) into Eq. (3.3.3), the temperature
gradient is expressed as:
E& 2 8(1 + ε s )ρ m a
sin (2 x / D ) − ⋅
dTm ∏ δ κ (1 + l / y0 ) p A2 (3.3.14)
= .
dx B + D ⋅ cos(2 x / D )
Obviously, the temperature gradient reduces as E& 2 increases. Part of the acoustic
power is “consumed” to establish the energy flow E& 2 and is no longer available to
generate a temperature profile. Integration of the temperature gradient Eq. (3.3.14)
along the stack plate yields the temperature distribution:
If D + B cos(2 x / D ) ≥ 0
x stack + Lstack dTm − a2   2 x 
∆Tm ( x) = ∫ dx = ln  B + D ⋅ cos 
x stack dx 2c p (1 + l / y0 )   D 
x stack + l stack
E& 8(1 + ε s )ρ m a D/2 B 2 − D 2 sin (2 x / D )
− 2 ⋅ ⋅ arctg
∏ δ κ (1 + l / y0 ) p A2 D + B cos(2 x / D )
(3.3.15a)
B2 − D2 x stack

If D + B cos(2 x / D ) < 0
Standing-wave systems 43

− a2
ln[B + D ⋅ cos(2 x / D )]
x dTm
∆Tm ( x) = ∫ dx =
xstart dx 2c p (1 + l / y0 )
x stack + l stack
E& 8(1 + ε s )ρ m a D/2  B 2 − D 2 sin (2 x / D ) 
− 2 ⋅ ⋅ ⋅  π + arctg .
∏ δ κ (1 + l / y0 ) p A
2
B2 − D2 
 D + B cos(2 x / D ) 
x stack

(3.3.15b)
Here, lstack is the stack length. The influence of different total energy flows on the
system performance is investigated, for the example case of table 3.1.I, again at
drive ratio of 1% and 10%. In this section, the influence of stack plate spacing is
not further investigated. The stack plate spacing is fixed as four times the fluid’s
thermal penetration depth, i.e. y 0 / δ k = 2 and so is the thickness of the solid stack
plate, i.e. l / δ s = 2 .
In order to choose a realistic value, the energy flow is related to the maximum
conductive heat flow that may be present in the stack for the case of zero-total-
energy-flow. If we denote the reference value as “Max”, the energy flow values
used in the following computation is set to 0, 25%, 50% and 75% of “Max”.
 dT 
Max = ∏( y 0 K + lK s ) m  (3.3.16)
 dx  max
in the case of E& 2 = 0 , at the stack positions where the local temperature gradients
are maximum.
In the case of 10% drive ratio and y 0 / δ k = 2 , this Max value can be obtained
from corresponding computation of the above “zero total energy flow” part, (they
are obtained at positions near the closed end of the tube) which is 1.88 W for this
tube geometry and diameter. In the case of 1% drive ratio, the Max value is 0.1684
W. According to the computation in the foregoing section, Max=1.88 W in the case
of drive ratio 10%, and Max=0.1684 W for the case of 1%. The plots are shown
below.
Similar to Figs. 3.3.6 and 3.3.8: the local temperature gradient increases gradually
and reaches its peak value at some position near the closed end, then drops
dramatically as the stack moving towards the close end. The temperature gradient
deteriorates only minor in the left part of the tube (0-0.20 m), but there is a fairly
dramatic drop in the region at x=0.225 m and further. The reason is that due to the
additional load, there must be a temperature drop at the end of the stack causing a
large negative dT/dx at the last 1% of the stack end. When the total energy flow
increases, the peak values drop and the corresponding stack position is moving
away from the closed end, the pressure anti-node.
44 Chapter 3

12000

10000
E2=0.0
8000 E2=Max*25%
6000 E2=Max*50%
E2=Max*75%
4000
dTm/dx (K/m)

2000

-2000

-4000

-6000

-8000

-10000
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.10: Calculated temperature gradient in the stack at various stack


positions and different total energy flows, drive ratio=10%.

E2=0.0
1200
E2=Max*25%
1000 E2=Max*50%
800
E2=Max*75%

600

400
dTm/dx (K/m)

200

-200

-400

-600

-800

-1000
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

Figure 3.3.11: Calculated temperature gradient in the stack at various stack


positions and different total energy flows, drive ratio=1%.
Standing-wave systems 45

400

Temperature difference over the stack (K)


350

300 E2=0.0
E2=Max*25%
250 E2=Max*50%
E2=Max*75%
200

150

100

50

0
0.00 0.05 0.10 0.15 0.20
Stack position (m)

Figure 3.3.12: Calculated temperature difference over the stack at different stack
positions and different total energy flows, drive ratio=10%.

60 E2=0.0
E2=Max*25%
Temperature difference over the stack (K)

50 E2=Max*50%
E2=Max*75%
40

30

20

10

-10
0.00 0.05 0.10 0.15 0.20
Stack position (m)

Figure 3.3.13: Calculated temperature difference over the stack at different stack
position and different total energy flows, drive ratio=1%.
46 Chapter 3

In Fig. 3.3.11, at lower drive ratios, the effect can be seen more profoundly. There
is a large effect of the energy load at lower pressure amplitudes. It is because the
dT/dx|max approaches dT/dx|critical at high pressure amplitudes (see Fig.3.3.6 curves
of y 0 / δ k = 2 and critical temperature gradient), but dT/dx|max approximates
1/10·dT/dx|critical at low pressure amplitudes (see Fig.3.3.8 curves of y 0 / δ k = 2 and
critical temperature gradient). But the energy grows at the square of pressure
amplitude, meaning that the energy of 10% drive ratio case is 100 times larger than
that of the 1% case.
A similar trend can be found in the figures of temperature difference over the stack,
Fig. 3.3.12 and 3.3.13. Again, the drive ratio has a big influence on the
performance of the system.
As shown in Fig.3.3.10 and 3.3.11, the local temperature gradient is negative at the
closed end of the resonator tube. That is due to the E& 2 value we assigned. The
substitution of Eq. (3.2.7) and (3.3.5) into (3.3.3) yields:

E2 =
& − ∏ δ κ ρ m c p u1s ( ) 2
T β
+ ∏ δ κ m p1s u1s − ∏( y0 K + lK s ) m
dTm 1 dT
4(1 + ε s )ω dx 4 1+ εs dx
(3.3.17)
In the presented graphs, the energy flow was related to the reference value “Max”
as
 dT 
E& 2 = c ⋅ ∏( y0 K + lK s ) m  = c ⋅ [∏( y0 K + lK s )] m
dT
 dx  max, E& 2 = 0 dx max, E& 2 = 0

(3.3.18)
in which c was set to 0, 0.25, 0.5, and 0.75.
At the closed end of the resonator tube, the particle velocity is zero, i.e. u1s = 0 ,
and thus the first two terms on the right hand side in Eq. (3.3.17) are zero.
Then, the gradient at the closed end can be expressed by substituting Eq. (3.3.18)
into (3.3.17) as
dTm dTm
= −c , (3.3.20)
dx closeend dx max, E& 2 = 0

which becomes negative for positive values of c .


A similar explanation can be made for the negative values at x=0 in Fig.3.3.10 and
3.3.11. (which is hardly visible, because of the extended scale.)

3.3.2 General analysis with viscosity included


In this section, many of the assumptions made before are eliminated. The
assumption that a standing-wave is sustained inside the resonator tube without
disturbance from the presence of the stack is maintained. The sound wave is
Standing-wave systems 47

described by Eq. (3.3.1) and (3.3.2) and now viscosity is included. The temperature
dependence of the various parameters is also taken into account. Ideal gas property
is assumed, i.e. Tm β = 1 . The ratio εs is neglected, i.e. εs≈0. As shown in the
definition of εs by Eq.(2.1.34), neglecting εs is realistic because the solid material
heat capacitance is much larger than that of working gas.
Substitution of Eq. (3.3.1) for the pressure wave and (3.3.2) for the velocity into
the general total energy equation (2.1.72) yields
~
E& 2  l  y p A2 2x  ( fκ − fν ) 
+ 1 +  0 sin( ) Im  ~ 
dTm ∏  y0  4ωρ m D D  (1 − fν )(1 + σ ) 
= 2
dx 1 + l 
 y0  y0 c p p A2  x 
2 ~
 ~ ( fκ − fν ) 
 ⋅  cos( )  ⋅ Im  fν +  − ( y0 K + lK s )
1 − fν
2
2ω 3 ρ m (1 − σ ) D 2  D   (1 + σ ) 
(3.3.21)
This equation is solved numerically using the following procedure:
1. The stack plate is divided into elementary cells
2. The hot end of the stack is sustained at 300K
3. Computation starts from the hot end of the stack
4. For the first computational cell at the hot end of the stack, the temperature
gradient is calculated with Eq. (3.3.21) based on gas properties
5. A new temperature for the immediate next computation cell is obtained via the
temperature gradient.
6. This computation is iterated to the other end of the stack.
The configuration that is considered is the same as discussed before and
summarized in table 3.3.I a and b with a drive ratio of 10%. Only now, all
parameters are considered temperature dependent. The thermal conductivity of
helium is
W
K = 0.0038T 0.65 . (3.3.22)
Km
And the viscosity can be expressed as
µ = 0.52T 0.64 × 10−6 Pa ⋅ s / K 0.64 . (3.3.23)

Zero total energy flow


The temperature difference across the stack ∆Tstack is calculated for different
positions of the stack inside the tube. In contrast to the previous zero-viscosity case,
we now will also investigate the effect of the operating frequency, or in other
words of the tube length. In addition to the 1 kHz considered so far, we now will
evaluate temperature profiles for 2 kHz and 3 kHz as well. When the operating
48 Chapter 3

frequency increases to 2 and 3 kHz from 1 kHz, the resonator tube length has to be
decreased to 1/2 and 1/3 of the length of 1 kHz correspondingly.

300

250

200
∆Tstack (K)

150

100 1 kHz
2 kHz
3 kHz
50

0
0.00 0.05 0.10 0.15 0.20 0.25
Stack position (m)

Figure 3.3.14: Temperature difference over the stack at various stack positions and
zero total energy flow, drive ratio=10%.

The temperature difference over the stack in these three computational cases is
shown in Fig. 3.3.14.
In the computations, the following parameters are fixed:
y 0 / δ k = 2.0 and l / δ s = 2.0 ; stack length/wave length=1/20; resonator tube
diameter/wave length=1/20; the resonator tube length/wave length=1/4 (a quarter
wave-length cooler). The temperature of the hot end of the stack is fixed at 300K in
all the computation cases.
Apparently, at higher operating frequency, the attainable temperature difference
decreases.
If the temperature difference over the stack is divided by the temperature of the hot
end of the stack, and the stack position is divided by the length of resonator tube,
Thot − Tcold X
T* = and X * = stack (3.3.24)
Thot Ltube
then the dimensionless plots are shown in Fig.3.3.15. The maximum temperature
difference across the stack decreases when the refrigerator is smaller or the
operating frequency is higher.
Standing-wave systems 49

1.0

0.8
1 kHz
2 kHz
3 kHz
0.6
∆Tstack/THot

0.4

0.2

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
*
Dimensionless stack position x =xstack/resonator length

Figure.3.3.15: Dimensionless temperature difference over the stack at different


stack positions along x with zero total energy flow, drive ratio=10%.

However, in the dimensionless figure, most part of the three curves overlap,
whereas their peak parts diverge. Perhaps, it can be explained as follows:
1. According to Eq. (3.3.21), using E& 2 = 0 , the temperature gradient can be
expressed as
~
p A2 2x 1~ ( fκ − fν ) 
sin( ) Im  fν + 
dTm 4ωρ m D D  (1 + ε s )(1 + σ ) 
=
dx  ~ f 
2  ( fκ − fν )(1 + ε s ν ) 
1 cp p 
2
x  ~ fκ  lK 
⋅ 2A  cos( )  ⋅ Im  fν +  −  K + s 
ω 2ω ρ m (1 − σ ) D 
2
D   (1 + ε s )(1 + σ )   y0 
 
(3.3.25)
The imaginary parts of the complex combination of Rott’s functions are always
negative in sign (to be shown in the next section). Rewriting it as:
dTm − C1 C1
= = ,
dx − 1 ⋅ C − ( K + lK s ) C2 + ( K + lK s ) (3.3.26)
ω 2
y0 ω y0
where C1 and C2 are positive functions related to the numerator and denominator
in Eq. (3.3.25). They remain constant when changing the frequency.
2. The length of the stack scales with 1 / ω .
50 Chapter 3

3. Using the results from 1 and 2, the total temperature difference across the stack
scales with
C1 1 C1
∆Tm ∝ ⋅ = .
+ (K + s ) ω
C2 lK lK
C2 + ω ( K + s ) (3.3.27)
ω y0 y0
Thus, the temperature difference over the stack becomes less when the frequency
increases. It is because the conductive loss becomes larger in the proportion of the
total energy flow while scaling down.

Non-Zero total energy flow


An energy flow is assumed to be present along the stack that is related to the
maximum conductive flow in case of zero energy flow—“Max”. As we did before,
the energy flow is set to 0, 0.25, 0.5 and 0.75 of this “Max” value. The resulting
temperature difference over that stack for 1 kHz operating frequency is plotted in
Fig.3.3.16. Here, the reference value Max=1.732 W. From Fig. 3.3.16, we can see
that the temperature difference over the stack drops if an energy load is applied to
the stack at the same stack position. The temperature drop becomes obvious when
the stack is closer to the close end. It is the same trend observed in the zero-
viscosity and non-zero total energy flow computation in section 3.3.1.

300

250 E2=0.0
E2=Max*25%
E2=Max*50%
200 E2=Max*75%
∆Tstack (K)

150

100

50

0
0.00 0.05 0.10 0.15 0.20 0.25
Stack position (m)

Figure 3.3.16: Calculated temperature difference over the stack at different stack
positions along x with different total energy flows, drive ratio=10%.
Standing-wave systems 51

It is apparent that the temperature span across the stack has a dependency on the
total energy flow along the stack. To see this effect, a plot of the temperature
difference over the stack at different total energy flows is shown in Fig. 3.3.17. In
this plot, the computation is made for the cases that the stack is fixed at positions
near the closed end (xstack=0.153 m), and a position somewhere between the closed
end and the pressure node (xstack=0.076 m).

120

xstack=0.153 m
100 xstack=0.076 m

80
∆Tstack (K)

60

40

20

0
0 10 20 30 40 50 60 70 80 90 100
Total energy flow E2 (W)

Figure 3.3.17: Temperature difference over the stack at a fixed stack position,
where the maximum temperature difference over the stack is achieved, with various
total energy flows along the stack, drive ratio=10%.

Fig.3.3.17 shows that the temperature span across the stack decreases rapidly at
increasing total energy flow. At some point, the maximum temperature span across
the stack becomes zero, which means that no temperature drop at the cold end is
available anymore with the specific amount of total energy flow along the stack.
The curve for the stack position near the closed end decreases rapidly, while the
other one for the position between the closed end and the pressure node decrease
slowly.

3.3.3 General analysis of “TAC”


Introduction
A “thermoacoustic couple” (TAC) refers to the simplest class of thermoacoustic
devices with a stack consisting of short plates without heat exchangers at any end,
and was introduced by Wheatley et al [41]. The analysis of a TAC is helpful in
52 Chapter 3

better understanding the basic thermoacoustic heat transport and also the coupling
between stack and its adjacent heat exchangers. Much work is done on TACs in
experiments and numerical simulation. Wheatley et al [41] measured the
temperature difference developed across a TAC as a function of its position in an
acoustic standing wave, with low drive ratio. They also developed a theoretical
expression to predict the temperature difference, stating from the point that the heat
transferred via entropy flow in the gas is returned by diffusive conduction. It was
assumed that the heat transfer between the couple and its surroundings was
negligible, because thermal insulation was applied in their set-up. Later, Atchley et
al [65] used similar TACs and extended the measurements to higher drive ratios up
to 2%. By comparing the measurement results with calculations based on the
theoretical expression developed by Wheatley et al [41], it was found that the
agreement was good for drive ratios below approximately 0.4%. At higher drive
ratios, the temperature difference along the stack was increasingly overestimated.
In some later experimental investigations [66, 67], the predictions based on linear
theory could not match the measured values either. Large discrepancies of up to
300% were reported. These deviations from linear thermoacoustic theory attracted
much attention and a succession of works was conducted to explore the
attributions. As stated by Piccolo and Cannistraro [67], possible causes can be: 1
non-linear effects due to the presence of harmonics higher than the fundamental; 2
turbulence and vortex generation; 3 heat leak to the surroundings; 4 additional
thermal load to the cold end carried by acoustic streaming caused by heat generated
in viscous losses along the resonator wall; 5 heat flow in transverse direction; 6
heat transferred to the ends of the stack in axial direction due to imperfect thermal
isolation. Many numerical simulations [68, 69], in which perfect thermal isolation
from the surroundings was assumed, were done to evaluate the heat flow in the
transverse direction. In standard linear thermoacoustic theory, this heat flow is
neglected by stating that the solid wall and adjacent gas have the same time-
averaged temperature. In the numerical simulation in reference [68], a discrepancy
up to 25% could be explained, which is only a small part of the large practical
discrepancy between linear theory and experimental data mentioned above. Paul
Aben experimentally made a detailed study about these nonlinear effects on a set-
up similar to “TAC” in his PhD work [84]. He foused on the vortex shedding at the
end of a parallel-plate stack; dissipation at the ends of a stack due to the sudden
change in cross section; transition to turbulence in-between plates; and streamings
taking place in a standing-wave device. His PIV visualization on the vorticity
pattern, and jet streaming, natural convection also by PIV technique has showed a
complicated situation at the ends of the stack.
Standing-wave systems 53

In the next section, an experimental apparatus similar to the so-called “TAC” was
used, as shown in Fig. 3.3.18. It is driven by a loudspeaker, which generates
acoustic power into the resonator tube at one end. The other end of the resonator is
closed by an end plug. A stack is placed inside the resonator tube and its position
can be adjusted by two steel tubes. The stack is housed in a stack cage and consists
of many parallel plates separated by fishing lines. There are no heat exchangers at
both ends of the stack. An analytical method for computing the temperature
distribution along the stack is proposed. Although no thermal insulation for the
system was employed, heat exchange with the environment is neglected in the
proposed analytical model.

Stack

Loudspeaker Resonator tube End plug

Figure 3.3.18: Schematic drawing of the apparatus similar to “TAC” and used for
the experiment.

Theory developed for the apparatus


The acoustic field inside the resonator tube is described by
p1 = p A cos( x / D ) and u1 = i ( p A / ρ m a )sin( x / D ) (3.3.28)
The acoustic field inside the stack is described by
p1 = p A cos( x / D ) and u1 = i (1 + l / y0 )( p A / ρ m a )sin( x / D ) (3.3.29)
where p A is the maximum acoustic amplitude.
Ideal gas is assumed, i.e. Tm β = 1 , and ε s is neglected, i.e. ε s = 0 . Substitution of
Eq. (3.3.29) into (2.1.72), in which the pressure gradient is replaced by the relation
for the acoustic velocity, in that case leads to
 Ares  Π y0 − p A2 ~
   2x   fk − fv 
E2 =
& ⋅ ⋅ sin   Im 1 − ~ 
 A  4ωρ D  D   (1 − fν )(1 + σ ) 
 gas  m

 A / A  2 Πy c  ~ f − f v   dTm
~
p2 x
+  res gas  ⋅ 3 0 p ⋅ 2A ⋅ sin 2   ⋅ Im  f v + k  ⋅
 1 − fν  2ω ρ m (1 − σ ) D  D   1 + σ   dx
 A  dT
− Π y0 K + solid K s ⋅ m , (3.3.30)
 ∏  dx
54 Chapter 3

where Ares is the cross sectional area of resonator tube and Agas is the cross sectional
area of gas in the stack. Asolid is the total cross sectional area of all solid material,
including stack cage, stack plates, and fishing lines, is given by
Asolid = ∏ l + Acage + N fish ⋅ A fish , (3.3.31)
where Acage and Afish are the cross sectional area of the stack cage and single fishing
line, respectively. П is the total effective plate length in cross sectional view.
Hence, the temperature gradient is
~
dTm  E& 2  Ares  y0 p A2  2x   fk − fv  
= + ⋅ ⋅ sin   Im − ~  ⋅
dx  Π  Agas  4ωρ m D  D   (1 − fν )(1 + σ )  
 A / A  2  ~ fk − fv 
~
2 x 
2
y0c p p
 
 2ω 3 ρ (1 − σ ) ⋅ D 2 ⋅ sin  D  ⋅ Im  f v + 1 + σ 
res gas A

 1 − fν  m  
−1
 A 
−  y0 K + solid y0 K s  . (3.3.32)
 Agas 
 
In order to use Eq. (3.3.32) to obtain the temperature gradient, it is necessary to
find the total energy flow along the stack E& 2 . In the present situation of
experimental apparatus, it is difficult to obtain the exact number for E& 2 . A
numerical simulation method could be time expensive and therefore is not an
option for the present work. In order to obtain an estimation for E& 2 by using the
available measurement devices, we assumed that the energy exchange with the
environment along the whole resonator tube is negligible. Since the thermometers
were installed at the centers of both ends of the stack, they are far from the wall of
the resonator tube, the temperatures are weakly influenced by the environment.
Therefore, the assumption is acceptable. Thus the total energy flow into the
resonator tube is the total acoustic power from the speaker. In the experiments, the
total acoustic power from the speaker can be calculated by the measured acoustic
pressure and volume velocity. Thus, the total energy flow E& 2 along the stack
without any loss is given by:
E& 2 = Pacoustic = 0.5 ⋅ p 1speaker ⋅ U1speaker , (3.3.33)
where Pacoustic is the total acoustic power from the speaker into the resonator tube. It
is calculated by the acoustic pressure p1speaker from microphone and volume velocity
U1speaker at the interface between the resonator tube and speaker diaphragm. U1speaker
is obtained by using the measurement from the accelerometer mounted on the
speaker diaphragm. As will shown later, here p1speaker and U1speaker are in phase
guaranteed by equipment measurement. Therefore, the phase terms are not needed
in Eq.(3.3.33). As a comparison, that total energy flow E& 2 is zero, the case of
Standing-wave systems 55

E& 2 =0, (3.3.34)


is also used for the later computation. We name the computation by using Eq.
(3.3.33) as method 1 “M1”, and by Eq.(3.3.34) as method 2 “M2”, which were
used in the figures later.

Acoustic power losses in the “TAC”


The “TAC” system is divided by the stack into three sections, as shown in Fig.
3.3.19.

Stack Plug
Sound wave
1
2 3
Loss 1 Loss 2 Loss 3

Figure 3.3.19: Three loss sections of the resonator tube.

The three sections are:


zone 1, the resonant tube segment from the loudspeaker to the nearest end of the
stack;
zone 2, the stack zone;
zone 3, the resonant tube segment from the other end of the stack to the plug.
Losses in these three sections are computed by two methods in this work.
(1) By solving the wave equation in section 1 and 3:
The acoustic field can be determined by solving the acoustic wave equation in the
tube sections 1 and 3 by using the pressure and velocity amplitudes at the two ends
of the resonator tube. Using the wave equation (2.1.57) in chapter 2, and assuming
that the temperature gradients in sections 1 and 3 are zero, and ε s is zero.
The wave equation in section 1 and 3 becomes:
a2 d 2 p1
(1 + (γ − 1) f κ )p1 + 2 (1 − fν ) 2 = 0 . (3.3.35)
ω dx
The Rott’s functions are evaluated with the parameter of the resonator tube. The
general solution of Eq. (3.3.35) is
p1 = C1 ⋅ e ik1x + C 2 ⋅ e ik 2 x , (3.3.36)
where C1 and C 2 are complex constants, only depending on the boundary
conditions, and k 1, 2 are two complex square roots
56 Chapter 3

ω 1 + (γ − 1) f κ
k 1, 2 = ± . (3.3.37)
a 1 − fν
Define real wave number
kr = ω a , (3.3.38)
and complex number
1 + (γ − 1) f κ
χ + iξ = . (3.3.39)
1 − fν
Thus, the general solution can be rewritten as
p1 = C1 ⋅ e − k rξ x ⋅ e i k r χ x + C 2 ⋅ e k rξ x ⋅ e −ik r χ x . (3.3.40)
At x=0, where the speaker is located, the pressure and velocity are measured
p1 x = 0 = p1speaker and u1 x = 0 = u1speaker (3.3.41)
By these boundary conditions, the complex constants are obtained
1 ρ m a ⋅ u1speaker 1 
C1 =  p1speaker − ⋅ , (3.3.42a)
2 1 − fν χ + iξ 
1 ρ m a ⋅ u1speaker 1 
C2 =  p1speaker + ⋅ . (3.3.42b)
2 1 − fν χ + iξ 
Thus, by substitution of Eq. (3.3.42a) and (3.3.42b) into Eq. (3.3.40), the pressure
at any location in section 1 is known. Similarly, by using the boundary conditions
at the end plug:
p1 x = L = p1e and u1 x = L = u1e = 0 (3.3.43)
the pressure distribution in section 3 is obtainable by the same method.
By using Eq. (3.3.40), the pressures and velocities at both ends of the stack, i.e. xL
and xR in Fig.3.3.20, are obtainable, indicate them as p1L , u1L , p1R ,and u1R ,
respectively.

stack
p1L p1R
u1L u1R

xL xR

Figure 3.3.20: Acoustic pressures and velocities in stack section.

The losses in the three sections with area A are


A ⋅ p1speaker u1speaker − A ⋅ Re[p1L u
~ ],
1 1
Loss1 = 1L (3.3.44a)
2 2
Standing-wave systems 57

A ⋅ Re[p1L u
~ ] − 1 A ⋅ Re[p u
1 R 1R ] ,
1 ~
Loss 2 = 1L (3.3.44b)
2 2
Loss3 = A ⋅ Re[p1R u~ ].
1
1R (3.3.44c)
2
In the figures for the loss computation in the next section, curves using this method
are indicated by “(1)”.
(2) By using the equation in chapter 2
The acoustic power variation in an elementary distance dx is given by Eq. (2.2.12)
in chapter 2. With assumptions made before, which are Tm β = 1 and ε s neglected,
i.e. ε s = 0 , Eq.(2.2.12) becomes:
 
(γ − 1) p1 2
2
dW&2 ρ u
~ 2 Im(− fν ) + ρ a 2 Im(− fκ )
1  
= − ∏ y0ω
m 1

dx 2 
 1 − fν m

~ ~
1 ∏ y0 1 dTm  ( fκ − fν ) ~ 
+ Re  ~ p1 u1  .
2 (1 − σ ) Tm dx
(3.3.45)
 (1 − fν ) 
In sections 1 and 3, the temperature gradients are assumed close to zero. Therefore,
the acoustic power losses in sections 1 and 3 are computed with
 
(γ − 1)p1 2
2
dW&2 ρ m u1
~ 2 Im(− fν ) + ρ a 2 Im(− fκ ) .
1  
= − Aresω (3.3.46)
dx 2 
1 − fν
 
m

The Rott functions fν and fκ are evaluated with the parameters of the resonater tube.
In section 2, which contains the stack, the acoustic power losses is computed by
using Eq. (3.3.45). The temperature gradient is computed by using Eq. (3.3.32) and
(3.3.33). In the computation, the losses in three sections are numerically integrated
by using Eq. (3.3.45) or (3.3.46).
In the figures for the loss computation in the next section, curves using this method
are indicated by “(2)”.
58 Chapter 3

3.4 Experimental results

In the next sections, the model, used for theoretical analysis about scaling-down a
standing-wave thermoacoustic refrigerator, is validated by experiments. An
experimental set-up was built, consisting of a resonator tube with a movable plug at
one end, a loudspeaker connected to the resonator by flanges, and stacks with
plates spacings at 0.2mm, 0.4 mm, and 0.6 mm. The stack can also be displaced by
steel tubes attached on it. These steel tubes contain the wiring of thermometers
placed on the stack. Microphones, thermometers, accelerometer and lock-in
amplifiers are used for necessary data acquisition. This set-up does not have heat
exchangers at the ends of the stack. So it is similar to a thermoacoustic couple
(TAC). The main function of this experiment is to validate the model, instead of
achieving large temperature drop or cooling power.
The performances at various stack positions are measured. Three stacks spacing at
0.2mm, 0.4 mm and 0.6 mm are used. The effects of different drive ratios, 0.5%,
1% and 2%, are investigated.

3.4.1 Experimental set-up

A schematic illustration of the standing-wave TAC is shown in Fig.3.4.1 and


photographs are given in Fig. 3.4.2 and 3.4.3. The set-up consists of a loud-
speaker, a resonator tube and a parallel-plate stack installed in a thin walled
cylinder. The length of the resonator tube can be adjusted by means of the plug.

Microphone 1 Stack in the Microphone 2


Steel rod
stack holder
Sound wave
Loudspeaker Resonator tube Plug Steel tubes
with an
Thermometers
accelerometer

Figure 3.4.1: Illustration of the standing-wave resonator tube.

The loudspeaker is housed in a cylindrical plastic housing, which prevents acoustic


energy from leaking out to the surroundings. It is a commercial moving-coil
loudspeaker, Dynaudio loudspeaker type D54 AF.
An accelerometer (Kistler 8614A1000M1, dev. is 1.67% in the frequency range of
300-600 Hz) is installed on the oscillatory diaphragm of the speaker. The voltage
and the current to drive the speaker are also measured with lock-in amplifiers (PAR
Standing-wave systems 59

5210). The resonator tube is made of transparent perspex, and has an inner
diameter of 25.7 mm and length of 546 mm. The loudspeaker diaphragm was
adapted, so that acoustic energy is radiated into the resonator tube. The speaker is
connected to one end of the resonator tube by flanges. The other end of the
resonator tube is closed with a plug. The plug can be displaced by pulling a steel
rod forth or backward. There are two locations for measuring the acoustic pressure.
One microphone (microphone 2, ENDEVCO 8510B-5, the uncertainty is ±0.2%
Full scale, ±69.6 Pa) is flush mounted at the end of the plug, facing the stack. The
other one (microphone 1, ENDEVCO 8510B-2, the uncertainty is ±0.2% Full
scale, ±27.6 Pa) is flush mounted at the wall of the resonator tube, close to the
speaker diaphragm. So the acoustic input can be measured. The pressure signals of
the two microphones are measured with lock-in amplifiers (PAR 5210). A stack of
parallel plates is housed in a thin walled cylinder. The stack position can be
changed by pulling steel tubes, which are attached to the stack holder. The
temperatures at both ends of the stack are measured by two resistance
thermometers (Jumo Pt 1000), and the resistance can be read from the “LakeShore”
218 temperature monitor. The working gas in the system is normal ambient air at
fixed pressure. There are no heat exchangers installed at the ends of the stack.
A schematic diagraph of the whole set-up is given in Fig. 3.4.2a. The measurement
devices are labeled with numbers in Fig. 3.4.2b and listed as follows:
1. 15 MHz Function/Arbitrary waveform generator, Hewlett Packard 33120A
2. System DC power supply, Agilent 6614C 0-100V/0-0.5A
3. System DC power supply, Agilent 6614C 0-100V/0-0.5A
4. Power supply, Delta Elektronika E015-2
5. Lakeshore 218 temperature monitor
6. 224 programmable current source, Keithley
7. speaker amplifier
8. 2*150 watts linear precision power AMP, Dynacord L300
9. Model 5210 Lock-in Amplifier 0.5Hz-120kHz, EG&G Princeton Applied
Research
10. Two channel Digital Real-Time Oscilloscope, Tektronix TDS210

The photo of the part of standing-wave system is given in Fig. 3.4.3.


60 Chapter 3

PC

Lock-in Amplifier

Lock-in Amplifier

Lock-in Amplifier

Lock-in Amplifier

Lock-in Amplifier
Lakeshore
218
temperature
monitor

T
p I V a p

Microphone 1 Microphone 2
Sound wave
Plug
Thermometers
Power amplifier

Wave form
generator

Figure 3.4.2a: Schematic diagraph of the experimental setup and measurement


devices.
Standing-wave systems 61

8
2,3 1

9 7 6 5 4

10

Figure 3.4.2b: Photo of the experimental setup: on the table, resonator and
loudspeaker set-up; on the left are electronics rack with lock-in amplifiers and
other meters.
Chapter 3

Figure3.4.3: Photo of the resonator and loudspeaker.


Microphone1
Stack in the
Microphone 2
stack cage
Resonator tube
Plug
Thermometers
Loudspeaker with an
accelerometer inside
62
Standing-wave systems 63

The stacks
The parallel-plate stacks consist of parallel plates made from Mylar material,
spaced by fishing line spacers glued between the plates. In the experiments, the
three spacing between the plates are 0.2 mm, 0.4 mm and 0.6 mm respectively. The
stacks are 21 mm long, with diameter of 22.7 mm and housed in a stack holder.
The stack holder is a 30 mm long, plastic cylinder, with a wall thickness of 1.5
mm. The cross section of the stack is shown in Fig. 3.4.4. The photographs of the
stacks are given in Figs. 3.4.5 and 3.4.6, and all relevant parameters are listed in
table 3.4.I.

Fishing line

Mylar plate

Figure 3.4.4 Illustration of the stacks.

Figure 3.4.5: Photo of the 0.6 mm, 0.4 mm and 0.2 mm-spacing stacks in the stack
holder (cross view).
64 Chapter 3

Figure 3.4.6: Photo of the 0.4 mm-spacing stack in the stack holder (longitudinal
view).

Stack
0.6 mm spacing 0.4 mm spacing 0.2 mm spacing
parameters
Half fluid layer
thickness y 0 0.3 mm 0.2 mm 0.1 mm

Half solid plate


0.05 mm 0.05 mm 0.05 mm
thickness l
Perimeter of the
1.08 m 1.55 m 2.65 m
stack plates Π
Isobaric
specific heat 1.01 E3 J/(kg·K) 1.01 E3 J/(kg·K) 1.01 E3 J/(kg·K)
cp
Diameter of the
25.7 mm 25.7 mm 25.7 mm
resonant tube
Area of the
7.1E-4 m² 7.1E-4 m² 7.1E-4 m²
speaker

Table3.4.I. Parameters used in computation of the experiment by utilizing the


model.
Standing-wave systems 65

Resonance frequency
Before conducting the thermoacoustic experiments on the stacks, introductory
experiments were performed to investigate the resonance frequency.
Two lock-in amplifiers are used to measure the acoustic pressure (from
microphone 1) and the acceleration of the diaphragm of the speaker. The resonance
frequency is defined as the frequency at which the phase difference between the
acoustic pressure and the acoustic volume velocity at the position of diaphragm of
the speaker is zero. Since the acceleration is ninety degrees out of phase with
respect to the local acoustic volume velocity, a ninety-degree phase difference
between the acceleration and the local acoustic pressure occurs at resonance. The
judgement of resonance is by reading the phase measurements of the acoustic
pressure and the acceleration from the two lock-in amplifiers. Besides that,
Lissajous curves are also a good visual means of observing resonance on an
oscilloscope. A ninety degree phase difference between the acceleration signal and
the acoustic pressure signal gives an undistorted ellipsoid on the oscilloscope
screen.

Time for temperatures to reach steady state


Prior to the actual experiments, the time required for thermal stabilization was
investigated. When the loudspeaker is switched on, the acoustic energy is coupled
into the resonator tube. As soon as the gas parcels around the stack begin to
oscillate, the thermal interaction between gas and stack plates starts to build up a
temperature gradient along these plates. There are no heat exchangers installed at
both ends of the stack. In this situation, both ends of the stack begin to warm up
and the temperature difference over the stack keeps increasing as well. After a
short period of time, this thermoacoustic process reaches its equilibrium. The
temperature difference between the two ends of the stack becomes stable. This time
needed for reaching the steady state, is determined experimentally before
conducting the series measurement, to make sure that all the data are taken in
steady state. To determine this time period, a trial measurement is taken. First, the
resonance frequency is found by the oscilloscope method mentioned above. The
stack is at a fixed position, and the loudspeaker loaded at that resonance frequency.
The moment when the loudspeaker is switched on is the “zero” point of time axis.
A plot of the time history of the temperatures at both ends of the stack is given in
Fig.3.4.7. It can be seen that the temperatures reach their steady states after
increasing for some time. Around 1000 seconds seen from Fig.3.4.7, the
temperatures stop increasing and stabilize at constant values. In the experiments of
this chapter, the start-up time is taken as 30 minutes for every measurement. It can
66 Chapter 3

be seen from Fig. 3.4.7 that 30 minutes, equal to 1800 seconds, is enough for a
specific operation to reach its steady state.

306

305

304

303
Temperature (K)

302

301

300

299

298

297

296

295
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Time (second)

Figure 3.4.7: Time development of the temperatures at two ends of the stack.

3.4.2 Measurements
Four kinds of measurements were conducted:
I. Fixed resonator tube length, resulting in a distance between end plug and
loudspeaker membrane of about 0.5m; fixed drive ratio (1% of the filling pressure)
and three stacks (0.2 mm, 0.4 mm and 0.6 mm spacing). The filling pressure was
always 1000 mbar.
All measurements were done with a fixed resonator tube length of 438 mm (the
distance from the flange surface, which is made from transparent perspex glued to
the end of resonator tube, and facing the loudspeaker, to the end plug) and the
pressure amplitude of the end plug was kept almost constant (the reading from the
lock-in amplifier of Microphone 2 remained almost constant). For a specific stack,
measurements were taken at a few stack positions. The stack position is specified
as the distance Lstack as indicated in Fig.3.4.8, which is the distance between the
diaphragm of the speaker and the nearest end of the stack.
II. Fixed resonator tube length to 438 mm; fixed drive ratio (2% of the filling
pressure); 0.6 mm-spacing stack and various stack positions.
All the measurements were done with a fixed resonator tube length 438 mm and
the stack of 0.6mm-spacing, and the pressure amplitude at the closed end was kept
almost constant at 2% of the filling pressure. The measurements were taken at a
Standing-wave systems 67

few positions of the stack. This experiment is to investigate the influence of the
driving pressure ratio on the performance of the system. The results are compared
with those of 1% drive ratio for 0.6 mm-spacing stack in measurement I.

Lstack
Sound wave
Loudspeaker Resonator tube Stack Plug

Figure 3.4.8: Illustration of the measurement position of the stack.

III. Fixed resonator tube length to 121 mm; different drive ratio (0.5% and 1% of
the filling pressure); 0.2 mm-spacing stack and various stack positions.
The measurements were done with a fixed resonator tube length 121 mm and the
stack of 0.2 mm-spacing, and the pressure amplitude at the closed end was kept
almost constant at 0.5% of the filling pressure. The measurements were taken at a
few positions of the stack. A second set of measurements was made at a drive ratio
of 1%. This experiment was also made to investigate the influence of the driving
pressure ratio on the performance of the system with different resonator tube length
and different stack from measurement II.
IV. Fixed resonator tube length to 121 mm; various drive ratio; 0.2 mm-spacing
stack and fixed stack position.
The measurements were done with a fixed resonator tube length 121 mm and the
stack of 0.2 mm-spacing, and the stack was fixed at Lstack=137 mm. The
measurements were taken at a few drive ratios. This experiment is to investigate
the influence of different acoustic pressure at the end plug surface, i.e. different
drive ratios, on the performance of the system.

For every measurement, the steps are as follows: place the stack at a specific
position and turn on the loudspeaker. Make a frequency sweep to search the
resonance frequency by the Lissajous criterion. Fix the loudspeaker resonance
frequency, and tune the amplitude of the acoustic wave generator such that the
acoustic pressure at the surface of the end plug is constant at around a ratio (0.5%,
1% or 2%) of filling pressure (keep the reading of microphone 2 constant). Make
the whole system run for 30 minutes. Record the readings from the lock-in
amplifiers for microphone 1, current and voltage flowing through the loudspeaker,
accelerometer, microphone 2 and temperatures of both ends of the stack from the
temperature monitor. Then, after this recording, the measurement procedure is
repeated for the next measurement point.
68 Chapter 3

Discussion of the results


I. From the measurement data, there are a lot of common features between the
cases of three stacks under drive ratio of 1%. Under the prerequisites of fixed
resonator tube and constant acoustic pressure at the surface of the end plug, with
the stack moving toward the end plug, the input current and voltage into the
loudspeaker decrease. That means the input total power to the whole system
decreases when the stack is moved closer and closer to the plug end of the
resonator tube, which is shown in Figs.3.4.9. The curves of 0.4 mm and 0.6 mm
stacks have similar values, but the curve of 0.2 mm stack gives much larger power.
The acoustic pressure in the vicinity of the loudspeaker does not vary much when
the stack moves towards the plug end of the resonator tube. The acceleration rate of
the diaphragm, however, decreases sharply. This shows that the total input acoustic
power into the resonator tube decreases, which is shown in Fig.3.4.10. Again, the
curves of 0.4 mm and 0.6 mm stacks have similar values, but the curve of 0.2 mm
stack shows much larger values. By comparison of corresponding curves in Fig.
3.4.9 and 3.4.10, the rates of electrical power converting to acoustic power for 0.4
mm and 0.6 mm stacks are around 10%, but the rate for 0.2 mm stack is much
lower, around 4%. According to the measurements of the cases of resonator tube
length 438 mm and 1% drive ratio, the resonance frequencies of 0.6 mm stack
varied in the range of 288 to 300 Hz; 288 to 300 Hz for 0.4 mm stack; and 259 to
300 Hz for 0.2 mm stack. The measurement also shows that the resonance
frequency slightly increases as the stack is moved towards the end plug.
In an ideal situation, a standing wave field exits inside the resonator tube. A pure
standing wave does not transfer energy. So, in the ideal cases, no acoustic energy is
needed. But, in reality, due to the viscous loss, thermodynamic loss and other
factors, continuous acoustic energy input is needed to compensate these losses and
thus maintain a standing wave. Figs. 3.4.10 shows that the acoustic energy input
decreases when the stack moves closer to the closed end. Since the closed end is a
velocity node, at a position closer to this end, the stack will exhibit lower viscosity
losses and thus the acoustic energy input required to maintain the standing wave
reduces.
The measured temperature difference over the stack ∆Tstack varies against stack
positions are given in Fig. 3.4.11. It shows that there is no much difference
between 0.6 mm stack and 0.4 mm stack. The temperature difference over the stack
∆Tstack of 0.2 mm stack is much lower than those of 0.4 mm and 0.6 mm stacks.
Among the three stacks, the 0.4 mm stack has the best performance. When the
stack moves to the end plug from the middle of the resonator tube, ∆Tstack increases
till a maximum, then decreases as the stack is close to the end plug.
Standing-wave systems 69

Total power into the speaker (W)


0.6 mm stack
5
0.4 mm stack
0.2 mm stack
4

0
0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.9: Total power input as a function of the stack position: 0.6 mm, 0.4 mm
and 0.2 mm spacing stack, resonator tube length 438 mm and 1% of drive ratio.

0.30
Total acoustic power into the resonator tube (W)

0.25
0.6mm stack
0.4mm stack
0.20 0.2mm stack

0.15

0.10

0.05

0.00
0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.10: Total input acoustic power as a function of the stack position: 0.6
mm, 0.4 mm and 0.2 mm spacing stack, resonator tube length 438 mm and 1% of
drive ratio.
70 Chapter 3

7
0.6mm stack
6 0.4mm stack
5 0.2mm stack

3
∆Tstack (K)

0
0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
-1
Stack position (m)
-2

-3

-4

Figure 3.4.11: ∆Tstack as a function of the stack position: 0.6 mm, 0.4 mm and 0.2
mm spacing stack, resonator tube length 438 mm and 1% of drive ratio.

II. This 0.6 mm-spacing stack with 2% drive ratio case has similar features as those
at 1% drive ratio. For both drive ratios, under the prerequisites of fixed resonator
tube and constant acoustic pressure at the surface of the end plug, with the stack
moving toward the end plug, the input current and voltage through the loudspeaker
decrease. That means the total electrical power to the whole system decreases when
the stack moving closer to the close end of the resonator tube, which is shown in
Fig.3.4.12. The 2% drive ratio case consumed more than proportional electrical
power than that of 1% case.
For each case, the acoustic pressure of microphone 1 does not vary much when the
stack is moved towards the closed end of the resonator tube. The acceleration rate
of the diaphragm decreases sharply. This shows that the total input acoustic power
into the resonator tube decreases, which is shown in Fig.3.4.13. The resonance
frequency becomes higher as the stack is moved towards the end plug. The
comparison of Fig.3.4.12 and 3.4.13 shows that both drive ratio cases have almost
the same conversion rate, around 10%, from electrical power to acoustic power.
Also, Figs. 3.4.12 and 3.4.13 show that the values of 2% drive ratio are nearly four
times the corresponding ones of 1% drive ratio.
The temperature difference over the stack ∆Tstack of two drive ratios is given in
Fig.3.4.14. When the stack is placed away from the end plug, less than 0.39 seen
from the figure, the difference of ∆Tstack between both drive ratios is not so large.
When the stack is closer to the end plug, say stack position larger than 0.39, the
Standing-wave systems 71

∆Tstack of the 2% case becomes much larger than that of 1% case. When the stack
position is further than 0.43, ∆Tstack of 2% is more than twice of that of 1% case.

1.6

1.4

Total power into the speaker (W)


1.2
drive ratio 1%
drive ratio 2%
1.0

0.8

0.6

0.4

0.2

0.0
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.12: Total power input as a function of the stack position: 0.6 mm
spacing stack, resonator tube length 438 mm, 2% and 1% of drive ratio.
Total acoustic power into the resonator tube (W)

0.18

0.16 drive ratio 1%


drive ratio 2%
0.14

0.12

0.10

0.08

0.06

0.04

0.02

0.00
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.13: Total input acoustic power as a function of the stack position: 0.6
mm spacing stack, resonator tube length 438 mm, 2% and 1% of drive ratio.
72 Chapter 3

14

12
drive ratio 1%
drive ratio 2%
10

8
∆Tstack (K)

0
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.14: ∆Tstack as a function of the stack position: 0.6 mm spacing stack,
resonator tube length 438 mm, 2% and 1% of drive ratio.

III. Similar comparison to measurement II was carried out for a different resonator
tube length of 121 mm and 0.2 mm stack. Two drive ratios 0.5% and 1% cases
show similar trends. For both drive ratios, with the stack moving towards the end
plug, the input current and voltage through the loudspeaker decrease. That means
the total electrical power to the whole system decreases when the stack is moving
closer to the closed end of the resonator tube, which is shown in Fig.3.4.15. The
1% drive ratio case consumed much more electrical power than that of 0.5% case.
Similarly, Fig.3.4.16 shows that the total input acoustic power into the resonator
tube decreases as the stack moves to the closed end for both cases. The 1% drive
ratio case consumed much more acoustic power than that of 0.5% case.
The comparison of Fig.3.4.15 and 3.4.16 shows that the energy conversion rate
from electrical power to acoustic power in this case, around less than 1% to 2%, is
much lower than case II. In case II, the resonance frequencies varied from 288 to
300 Hz. Taking the average frequency of 295 Hz, the gas thermal penetration depth
is around 0.15 mm. For the 0.6 mm spacing stack, y0=0.3 mm=2δκ, which is in
normal situation of a standing-wave type. In the 0.2 mm spacing case III, the
resonance frequency varied from 687 Hz to 712 Hz. Taking the average value of
700 Hz, the gas thermal penetration depth is around 0.1 mm, which just equal to
the y0 of 0.2 mm spacing stack. It means this stack has the spacing of a
regenerator, but was inserted in a standing-wave type resonator tube.
Standing-wave systems 73

The temperature difference over the stack ∆Tstack of two drive ratios is given in
Fig.3.4.17. For the case of drive ratio 1%, as the stack move to the end plug, ∆Tstack
increases till a maximum and then decreases when the stack is close to the end
plug. The ∆Tstack of the 1% case is much larger than that of 0.5% case.

20

18
drive ratio 0.5%
Total power into the speaker (W)
16 drive ratio 1.0%
14

12

10

0
0.08 0.09 0.10 0.11 0.12 0.13 0.14 0.15 0.16
Stack position (m)

Figure 3.4.15: Total power input as a function of the stack position: 0.2 mm
spacing stack, resonator tube length 121 mm, 0.5% and 1% of drive ratio.

0.16
Total acoustic power into the resonator tube (W)

0.14
drive ratio 0.5%
0.12 drive ratio 1.0%

0.10

0.08

0.06

0.04

0.02

0.00
0.08 0.09 0.10 0.11 0.12 0.13 0.14 0.15 0.16

Stack position (m)

Figure 3.4.16: Total input acoustic power as a function of the stack position: 0.2
mm spacing stack, resonator tube length 121 mm, 0.5% and 1% of drive ratio.
74 Chapter 3

7
drive ratio 0.5%
drive ratio 1.0%
6

5
∆Tstack (K)

0
0.08 0.09 0.10 0.11 0.12 0.13 0.14 0.15 0.16
Stack position (m)

Figure 3.4.17: ∆Tstack as a function of the stack position: 0.2 mm spacing stack,
resonator tube length 121 mm, 0.5% and 1% of drive ratio.

2
(Drive ratio)
-5 -4 -4 -4 -4
0.0 5.0x10 1.0x10 1.5x10 2.0x10 2.5x10
10 10
Total power into the speaker (W)

8 8

6 6

4 4

2 2

0 0
5 6 6 6 6 6
0.0 5.0x10 1.0x10 1.5x10 2.0x10 2.5x10 3.0x10
2 2
(Acoustic amplitude) (Pa)

Figure 3.4.18: Total power input to the speaker as a function of the acoustic
pressure square at the end plug and drive ratio square: 0.2 mm spacing stack,
resonator tube length 121 mm.
Standing-wave systems 75

2
(Drive ratio)
-5 -4 -4 -4 -4
0.0 5.0x10 1.0x10 1.5x10 2.0x10 2.5x10
0.20 0.20

Total acoustic power into the resonator tube (W)


0.18 0.18

0.16 0.16

0.14 0.14

0.12 0.12

0.10 0.10

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0.00 0.00
5 6 6 6 6 6
0.0 5.0x10 1.0x10 1.5x10 2.0x10 2.5x10 3.0x10
2 2
(Acoustic amplitude) (Pa)

Figure 3.4.19: Total acoustic power into the resonator tube as a function of the
acoustic pressure square at the end plug and drive ratio square: 0.2 mm spacing
stack, resonator tube length 121 mm.

14

12

10
∆Tstack (K)

2
400 600 800 1000 1200 1400 1600 1800
Acoustic amplitude (pa)

Figure 3.4.20: ∆Tstack as a function of the acoustic pressure at the end plug: 0.2 mm
spacing stack, resonator tube length 121 mm.
76 Chapter 3

IV. When the acoustic pressure amplitude at the end plug surface increases, the
power consumed by the system also increases, as shown in Fig.3.4.18. Similar
trends are also shown for the total acoustic power into the resonator tube, as shown
in Fig. 3.4.19. Figs. 3.4.18 and 3.4.19 also show that the power flows are linear
functions of the acoustic pressure amplitude squared. The temperature difference
over the stack ∆Tstack at different acoustic pressure is given in Fig.3.4.20. As
expected, ∆Tstack increases almost linearly as the acoustic pressure at the end plug
surface increases.

3.4.3 Theoretical computation


The computation results by using section 3.3.3 are discussed below.
(a) Computation of ∆Tstack as a function of stack positions or acoustic pressure at
the end plug
Experimental data for the cases of resonator tube length of 438 mm, along with
computation results by using Eq.(3.3.33) and (3.3.32) as method 1 “M1”, and by
Eq.(3.3.34) and (3.3.32) as method 2 “M2”, are given from Fig. 3.4.21 to Fig.
3.4.24. Computations and measurement data for the cases of resonator tube length
of 121 mm are plotted from Fig. 3.4.25 to 3.4.27. In the computation, the tube
length is set to half the wave length λ = a / f , where a is the speed of sound at
normal atmospheric pressure and ambient temperature 300K, and f is the
corresponding measured operating resonance frequency.

16
15 computation M1
14 computation M2
13 experiment
12
11
10
9
∆Tstack (K)

8
7
6
5
4
3
2
1
0
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.21: Comparisons between computation and experiment of case with 0.6
mm spacing stack, 1% drive ratio and resonator tube length 438 mm.
Standing-wave systems 77

18

16
computation M1
14
computation M2
experiment
12

10
∆Tstack (K)
8

0
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.22: Comparisons between computation and experiment of case with 0.4
mm spacing stack, 1% drive ratio and resonator tube length 438 mm.

14

12
computation M1
10 computation M2
8
experiment

6
∆Tstack (K)

0
0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
-2
Stack position (m)
-4

-6

-8

Figure 3.4.23: Comparisons between computation and experiment of case with 0.2
mm spacing stack, 1% drive ratio and resonator tube length 438 mm.
78 Chapter 3

32
30
28 computation M1
26 computation M2
24 experiment
22
20
18
∆Tstack (K)

16
14
12
10
8
6
4
2
0
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.24: Comparisons between computation and experiment of case with 0.6
mm spacing stack, 2% drive ratio and resonator tube length 438 mm.

20
18 computation M1
16 computation M2
experiment
14
12
10
∆Tstack (K)

8
6
4
2
0
0.08 0.10 0.12 0.14 0.16
-2
-4
Stack position (m)

-6

Figure 3.4.25: Comparisons between computation and experiment of case with 0.2
mm spacing stack, 1% drive ratio and resonator tube length 121 mm.

As seen in Figs. 3.4.21 to 3.4.27, the computations based on the proposed


analytical method by using Eq.(3.3.33) have better performance than those using
method 2. In Figs.3.4.23, 3.4.25, 3.4.26 and 3.4.27, the computation results have
lower ∆Tstack than those of the experiment. In the computations, a larger E& 2 leads
Standing-wave systems 79

to a lower ∆Tstack. Therefore, a lower ∆Tstack than experimentally measured is due to


the overestimation of E& 2 by Eq. (3.3.33). That can be explained by the assumption
that no energy exchange with the environment takes place.

5 computation M1
computation M2
4
experiment
3
∆Tstack (K)

0
0.08 0.09 0.10 0.11 0.12 0.13 0.14 0.15
-1
Stack position (m)
-2

Figure 3.4.26: Comparisons between computation and experiment of case with 0.2
mm spacing stack, 0.5% drive ratio and resonator tube length 121 mm.

28
26
computation M1
24
computation M2
22 experiment
20
18
16
∆Tstack (K)

14
12
10
8
6
4
2
0
400 600 800 1000 1200 1400 1600 1800
Acoustic amplitude (pa)

Figure 3.4.27: Comparisons between computation and experiment of the case with
0.2 mm spacing stack, stack position fixed at 137 mm and resonator tube length
121 mm with various acoustic pressures at the end plug.
80 Chapter 3

(b) The acoustic power losses


The losses in the three sections of the resonator for all cases are given from Fig.
3.4.28 to 3.4.34. The losses computed by using “by solving the wave equation in
section 1 and 3” (see section 3.3.3) are indicated as “(1)”, and those by “by using
the equation in chapter 2” (see section 3.3.3) are indicated as “(2)”.
As expected, the loss in section 1 increases, and the loss in section 3 decreases as
the stack moves towards the end plug. As seen in the figures from Fig.3.4.28 to
3.4.33, the loss in section 2 decreases as the stack moves towards the end plug. In
Fig.3.4.34, all losses in three sections increase as the acoustic pressure on the end
plug surface increases. This can be explained by the stronger acoustic field leading
to higher thermoviscous losses. Among three sections, the losses in section 2
increase more in scale than the losses in section 1 and 3.
In these figures, two methods agree well with each other in section 1 and 3, but
there are discrepancies in section 2. The discrepancy in section 2 is large for the
cases with 0.2 mm-spacing stack, which are shown in Figs. 3.4.30, 3.4.32, 3.4.33
and 3.4.34. We consider the method (1) is more accurate than method (2).

0.030 loss before the stack (1)


loss before the stack (2)
loss in the stack (1)
0.025 loss in the stack (2)
loss behind the stack (1)
loss behind the stack (2)
0.020
Energy (W)

0.015

0.010

0.005

0.000
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.28: Comparisons of losses between computation methods (1) and (2) of
case with 0.6 mm spacing stack, 1% drive ratio and resonator tube length 438 mm.
Standing-wave systems 81

0.045
loss before the stack (1)
0.040 loss before the stack (2)
loss in the stack (1)
0.035 loss in the stack (2)
loss behind the stack (1)
0.030 loss behind the stack (2)

Energy (W)
0.025

0.020

0.015

0.010

0.005

0.000
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.29: Comparisons of losses between computation methods (1) and (2) of
case with 0.4 mm spacing stack, 1% drive ratio and resonator tube length 438 mm.

loss before the stack (1)


0.25 loss before the stack (2)
loss in the stack (1)
loss in the stack (2)
loss behind the stack (1)
0.20
loss behind the stack (2)

0.15
Energy (W)

0.10

0.05

0.00
0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.30: Comparisons of losses between computation methods (1) and (2) of
case with 0.2 mm spacing stack, 1% drive ratio and resonator tube length 438 mm.
82 Chapter 3

0.14
loss before the stack (1)
loss before the stack (2)
0.12
loss in the stack (1)
loss in the stack (2)
0.10 loss behind the stack (1)
loss behind the stack (2)
Energy (W)

0.08

0.06

0.04

0.02

0.00
0.30 0.35 0.40 0.45 0.50
Stack position (m)

Figure 3.4.31: Comparisons of losses between computation methods (1) and (2) of
case with 0.6 mm spacing stack, 2% drive ratio and resonator tube length 438 mm.

0.16 loss before the stack (1)


loss before the stack (2)
0.14 loss in the stack (1)
loss in the stack (2)
loss behind the stack (1)
0.12
loss behind the stack (2)

0.10
Energy (W)

0.08

0.06

0.04

0.02

0.00
0.08 0.09 0.10 0.11 0.12 0.13 0.14 0.15 0.16
Stack position (m)

Figure 3.4.32: Comparisons of losses between computation methods (1) and (2) of
case with 0.2 mm spacing stack, 1% drive ratio and resonator tube length 121 mm.
Standing-wave systems 83

loss before the stack (1)


0.04
loss before the stack (2)
loss in the stack (1)
loss in the stack (2)
loss behind the stack (1)
loss behind the stack (2)
0.03

Energy (W)
0.02

0.01

0.00
0.08 0.09 0.10 0.11 0.12 0.13 0.14
Stack position (m)

Figure 3.4.33: Comparisons of losses between computation methods (1) and (2) of
case with 0.2 mm spacing stack, 0.5% drive ratio and resonator tube length 121
mm.

0.18
loss before the stack (1)
0.16 loss before the stack (2)
loss in the stack (1)
0.14 loss in the stack (2)
loss behind the stack (1)
loss behind the stack (2)
0.12
Energy (W)

0.10

0.08

0.06

0.04

0.02

0.00
400 600 800 1000 1200 1400 1600 1800
Acoustic amplitude (pa)

Figure 3.4.34: Comparisons of losses between computation methods (1) and (2) of
case with 0.2 mm spacing stack, stack position fixed at 137 mm and resonator tube
length 121 mm as a function of the acoustic pressure at the end plug.
84 Chapter 3

3.4.4 Conclusions
An analytical method is presented for evaluating the temperature profile along a
stack in a standing-wave “thermoacoustic couple”. A thermoacoustic resonator
tube was built, and experiments with a “thermoacoustic couple” were made. These
experiments have been compared with the model. The trend of the temperature
differences over the stack between measurement and model is the same, however, a
precise agreement depends much on the correct estimation of the total energy flow
along the stack E& 2 . Better agreement about temperature differences along the stack
∆Tstack between experimental measurement and computation by using the proposed
method 1 than method 2 assuming E& 2 =0 was obtained. If the total energy flow
along the stack E& 2 is accurately given, the model discussed in section 3.3 is able to
describe and to explore the standing-wave system.
Chapter 4

Traveling-wave systems

4.1 Introduction

Although a traveling-wave system has many similarities with a standing-wave


system, it executes a different thermodynamic cycle compared to a standing-wave
system. The intermaterial spacing in the regenerator of a traveling-wave system is
much smaller than in the stack of a standing-wave system. The phasing between
acoustic pressure and volume velocity in a traveling-wave system is nealy in phase,
whereas a standing-wave type is nearly 90 degrees out of phase. This phasing
relation makes a traveling-wave system work in a Stirling cycle. A Stirling cycle
has no intrinsic irreversibilities so that a higher thermal efficiency occurs. Under
ideal conditions, the ideal efficiency of the Stirling cycle is equal to the Carnot
efficiency. This high efficiency makes a traveling-wave system more competitive
in applications than its counterpart, the standing-wave system, which involves an
intrinsically irreversible process.
Stirling machines originated from the invention of Robert Stirling in 1816. His
ingenious design went ahead of the development of thermodynamic theory.
Afterwards, much work has been done fruitfully based on his patent [55, 56].
In 1964, an important discovery by W. Beale [85] made an inspiration to a piston-
free Stirling machine. While teaching a course on thermal machinery at Ohio
University, he found that the piston was not necessary in a Stirling β-configuration,
once the machine started. Because the machine could continue to work only based
on gas springs with the correct phasing relationship. The first device without any
sliding seals was the Fluidyne engine, which used U-tube liquid pistons [57]. But
its performance is limited to low frequencies due to the large mass of the liquid
pistons. Ceperley discovered that a traveling acoustic wave propagating through a
differentially heated regenerator executes a thermodynamic cycle similar to the
Stirling cycle [58, 59]. As the wave travels up the temperature gradient through the
86 Chapter 4

regenerator, the wave is amplified and thermal energy converts into acoustical
energy. The opposite direction results in acoustical energy being used to pump heat.
Although Ceperley’s experimental engine was not able to amplify acoustic power,
his remarkable idea stimulated interest in using sound to build piston-free Stirling
type engines and refrigerators. Much later, Yazaki et al. first demonstrated a stack-
based traveling-wave amplifer [60]. Unfortunately, the engine had a low efficiency.
They realized that the low acoustic impedance of the working gas caused large
viscous losses due to high acoustic velocities. In 1999, Backhaus and Swift
presented a new configuration of a thermoacoustic Stirling heat engine [61] and
later, in 2000, they made a detailed description of it [34]. Their ingenious invention
solved the problem of low acoustic impedance, made a breakthrough in building a
thermoacoustic engine to achieve the efficiency of a conventional gasoline-
powered engine.
In the next section, a physical description of traveling-wave systems is given.
Section 4.3 presents an analytical model for a complete traveling-wave system of
refrigerator. After that, in section 4.4, the regenerator material is considered.
Experiments are performed on various regenerator materials in which also the
model developed in section 4.3 is validated. A full traveling-wave refrigerator
driven by a mechanical compressor was designed and built. This system is
described in section 4.5 along with experiments again aiming at the validation of
the model.

4.2 Physical description of traveling-wave systems

A simplified physical description about how a traveling-wave thermal process


develops is given in this section. A Stirling cycle for the refrigeration mode will be
presented. A schematic description of an ideal Stirling cycle is shown in Fig. 4.2.1.
From left to the right, it consists of a warm piston, a hot-end heat exchanger, a
regenerator, a cold-end heat exchanger and a cold piston.
In steady-state, starting at the moment when the warm piston is in its left most
position, and the cold piston is in contact with the cold-end exchanger, a Stirling
refrigeration cycle can be described in four steps:
1. From a to b. The warm piston moves to the right over some distance while
the cold piston is fixed. This isothermal process gives off heat flow Q& H to
the surroundings through the hot-end heat exchanger.
Traveling-wave systems 87

Warm Cold
piston Regenerator piston

a
TH TC

b p
b Q& H
Q& H 1
2
a
c c 4 TH
Q& C 3
Q& C d TC
V V2 V
d 1

Figure 4.2.1: Ideal Stirling cycle with the P-V diagram.

2. From b to c. Both pistons move to the right over the same distance to ensure
the total volume enclosed remains constant. The gas enclosed passes
through the regenerator. The gas enters the regenerator from the left end at
room temperature and leaves it at the right end with low temperature.
During this process, the heat is given off by the gas to the regenerator
material.
3. From c to d. The cold piston moves to the right while the warm piston
remains fixed. This isothermal process takes up heat Q& C from the
application via the cold-end heat exchanger. So, here cooling takes place.
4. From d to a. Both pistons move to the left over the same distance to ensure
the total volume enclosed remains constant. The gas enclosed passes
through the regenerator. The gas enters the regenerator from the right end at
low temperature and leaves it at the right end at high temperature. During
this process, the regenerator material is cooled by the cold gas. Then the
whole system returns to its initial state and is ready for the next cycle.
So netto heat is given off to a warm heat exchanger, and cooling takes place at the
cold heat exchanger. With this cooling power an application can be cooled.
The main part of a thermoacoustic traveling-wave refrigerator, which contains the
regenerator, retains all components of their mechanical equivalent, except for the
pistons. The removal of the moving pistons of a Stirling refrigerator is realized by
using an acoustic wave and a properly dimensionalized acoustic resonator.
88 Chapter 4

4.3 Modeling traveling-wave systems

Compliance
Ambient HX
Regenerator Feedback
Cold end HX inertance

Driver Resonator tube

Figure 4.3.1: A traveling-wave refrigerator driven by a loudspeaker.

In this section, a model for a traveling-wave system with a torus-shaped section is


discussed, as depicted in Fig. 4.3.1. The system of Fig. 4.3.1 employs a
thermoacoustic conventional configuration, using a large volume called
compliance to connect the feedback inertance and the ambient heat exchanger.
However, the author found that the compliance is not necessary for the
performance of a traveling-wave system. Therefore, in the section 4.5, a traveling-
wave refrigerator without a compliance was designed and used as the experimental
apparatus. For the present section, in the consideration of generality, the theoretical
model sticks to the conventional configuration.
In Fig.4.3.1, the refrigerator system is driven by a driver, feeding acoustic power
into the refrigerator part. A resonator tube is connected to the driver at one end and
to the torus-shaped section at the other end. The length of the resonator tube
depends on the wavelength of the working gas. The torus-shaped section contains a
feedback inertance tube, a compliance volume, and a regenerator which is
sandwiched between two heat exchangers. A traveling-wave system in such a
configuration, which was first developed by Backhaus and Swift in the mode of an
engine [34], has a favorable efficiency and thus is valuable for applications and
research of the present work.
Much work was focused on numerically modeling a traveling-wave system.
Among them, SAGE and DeltaE are the most extensively used tools for modeling
in the thermoacoustic community. Swift and his coworkers developed a computer
program DeltaE which integrates the acoustic wave equations, and can be used for
apparatus ranging from simple duct networks and resonators to thermoacoustic
prime movers, refrigerators, and combinations at low-pressure amplitudes [71].
SAGE is developed by D. Gedeon, and utilizes finite difference methods to
simulate the machine and predict the performances [72]. This commercially
accessible software provides guidance in thermoacoustic system design. W.Dai et
Traveling-wave systems 89

al. [73] developed computer codes to simulate the fluid field inside a
thermoacoustic system and make performance predictions. All of these computer
programs numerically simulate the acoustic fields and predict the performances. It
normally takes much time to obtain a converged result by using these programs.
Although these numerical programs are powerful in simulation and prediction of a
thermoacoustic system, they are expensive in computation time. The analytical
method presented in this work has the advantage of predicting the performance in a
fast manner. Therefore, it is a useful tool in parameter-optimization studies. This
analytical expression is also helpful in obtaining a universal conclusion for scaling
analysis of traveling-wave thermoacoustic systems.

Control
Q& H
volumes

TC

W& in E& 2 W& dn−st


W&up− st

TH

Q& C

Figure 4.3.2: Energy flow distribution of a torus traveling-wave refrigerator.

In the regenerator of the looped configuration of a traveling-wave system, the


energy flow arrangement, is shown in Fig.4.3.2. The dash-dot rectangular frame
indicates the part containing the regenerator and two heat exchangers, where the
main heat exchange between the environment and the system takes place. For the
global system, the energy balance is
W&in + Q& C = Q& H , (4.3.1)
where W&in is the acoustic input power into the refrigerator by the driver.
The total energy flow along the regenerator is given by
W&up − st = Q& H + E& 2 , (4.3.2)
W&dn − st = E& 2 + Q& C . (4.3.3)
&
In a regenerator, the total energy flow E2 is very small. For an ideal regenerator,
and ideal gas as working gas, the heat capacity of the regenerator material and the
good thermal contact between working gas and solid regenerator material maintain
temporally isothermal conditions, so the temperature oscillation T1 = 0 everywhere.
If the conduction of heat through the regenerator can be neglected, Eq. (2.1.68)
shows that E& 2 = 0 [38].
90 Chapter 4

Using Eq. (4.3.3), the cooling power is


Q& C = W&dn − st − E& 2 . (4.3.4)
To analyze the behavior of the acoustic power flows and total energy flow along
the regenerator, the acoustic field has to be known.
In order to develop an analytical model, here below all the sections of the
traveling-wave system, as shown in Fig. 4.3.1, will be analyzed using the thermo-
acoustic equations. First the general case for a section with temperature gradient
and pressure gradient will be treated. From that the expressions for all the segments
are derived.
The most general wave equation that describes the behavior of the acoustic
pressure p1 in an acoustical pipe section is given by Eq. (2.1.57),
 (γ − 1) fκ  ρ a 2 d  1 − fν dp1  a2 fκ − fν dTm dp1
1 + p1 + m 2   − β 2 =0.
 1+ εs  ω dx  ρ m dx  ω (1 − σ )(1 + ε s ) dx dx
The Rott’s functions f v , fκ and ε s are evaluated locally and with corresponding
geometry. In order to solve Eq. (2.1.57) for the pressure, the temperature or
temperature gradient must be known. For most of the components in Fig. 4.3.1, the
temperature is constant. However, in the regenerator, the temperature varies in the
axial direction x, and the parameter f v , fκ , viscous penetration depth δν , thermal
penetration depth δ κ , sound speed a , density ρ m , gas conductivity κ and
viscosity µ are functions of x.
First the material properties like viscosity etc. are written as functions of
temperature via substution of Eq. (2.1.16). Substitution of (2.1.16), (A.6), (2.1.38)
and (2.1.31) into wave equation (2.1.57), after a long calculation, leads to
bµ + 3
 − 
2
d p1  fκ − f v (bµ + 3) f v bµ + 1 Tm 2  β
+ 1 − − + ⋅
dx 2
(1 − σ )(1 + ε s ) 2 2 cosh [(1 + i ) y0 / δ v ]  1 − f v
2
 
dTm dp1 ω 2 1 + (γ − 1) fκ /(1 + ε s )
⋅ ⋅ + ⋅ ⋅ p1 = 0 . (4.3.5)
dx dx a 2 1 − fv
This is a second order differential equation. Like in chapter 2, the solution has the
following form
p1 = Ceαx . (4.3.6)
Substitution of Eq. (4.3.6) into (4.3.5) yields
bµ + 3
 − 
 fκ − f v (bµ + 3) f bµ + 1 Tm 2  β
α + 1 −
2
− v
+ ⋅
(1 − σ )(1 + ε s ) 2 2 cosh [(1 + i ) y0 / δ v ]  1 − f v
2
 
Traveling-wave systems 91

dTm ω 2 1 + (γ − 1) fκ /(1 + ε s )
⋅ ⋅α + 2 ⋅ =0. (4.3.7)
dx a 1 − fv
For simplicity of writing, we define
bµ + 3
 − 
 fκ − f v (bµ + 3) f bµ + 1 Tm 2  β dTm
G = 1 − − v
+
(1 − σ )(1 + ε s ) 2 2 cosh [(1 + i ) y0 / δ v ]  1 − f v dx
2
 
(4.3.8)
and
ω 2 1 + (γ − 1) fκ /(1 + ε s )
H= ⋅ . (4.3.9)
a2 1 − fv
Thus Eq. (4.3.7) can be rewritten in terms of α
α2 + G ⋅ α + H = 0 . (4.3.10)
Therefore, the roots of Eq. (4.3.10) are
α1, 2 =
1
2
(
− G ± G 2 − 4H . ) (4.3.11)

Thus the general solution of acoustic pressure oscillation is


p1 = C1 ⋅ eα1 x + C2 ⋅ eα 2 x , (4.3.12)
where the complex coefficients C1 and C2 are determined by the boundary
conditions.
The boundary conditions depend on the specific configuration of the refrigerator
and also depend on local parameters. Therefore, it is difficult to obtain a universal
expression of the solution, certainly not for a looped tube system with temperature
gradients in it. However, in principle the solution (4.3.12) can be used to
component-wise couple the tube sections and find a universal expression for the
wave in the complete resonator. This will be explained below.
In order to solve the equations, an assumption of zero temperature gradient on the
solid wall, i.e. dTm / dx = 0 , is made for all components except the regenerator.
The regenerator, using the concept of differentiation, is divided into many
differential elements to resolve the solution.
Zero temperature gradient components
For the zero-temperature-gradient components, the acoustic pressure equation is
much simplified via Eq. (4.3.12) where dTm / dx = 0 has been used ( G = 0 ) and
reduces to
p1 = C1 ⋅ eikx + C2 ⋅ e-ikx , (4.3.13)
where the complex wave number is defined as:
92 Chapter 4

ω 2 1 + (γ − 1) fκ /(1 + ε s )
k2 = ⋅ . (4.3.14)
a2 1 − fv
Thus, it can be obtained that
[
dp1 / dx = ik C1 ⋅ eikx − C2 ⋅ e-ikx . ] (4.3.15)
By using Eq. (2.2.6), the y-direction averaged velocity follows as
u1 =
i

dp1
(1 − fν ) .
ωρ m dx
The volume velocity can be written as
U1 = u1 A = −
(1 − fν )Ak [C ]
⋅ eikx − C2 ⋅ e- ikx . (4.3.16)
ωρ m 1

In most practical cases, the components are tubes, at ambient temperature without
an axial temperature gradient, with a radius much larger than the thermal and
viscous penetration depths at ambient temperature.
We consider sound propagating in the x direction in an ideal gas within a channel
with cross-sectional area A and perimeter Π , the hydraulic radius is defined as
rh = A / Π . (4.3.17)
Now that the hydraulic radius of a tube segment rh >> δ v and δ κ can be sustained,
so Rott’s functions f v ≈ 0 and fκ ≈ 0 .
The wave number can now be reduced to a real number from Eq. (4.3.14)
k =ω/a , (4.3.18)
where the sound speed is at the working temperature of these tubes, usually,
ambient temperature.
The pressure and volume velocity for every tube segment are reduced to
p1 = C1 ⋅ eikx + C2 ⋅ e-ikx , (4.3.19)

U1 = u1 A = −
A
aρ m
[C ⋅ e
1
ikx
− C2 ⋅ e -ikx .] (4.3.20)

To describe the complete system, it is necessary to compute the acoustic


impedance section-wise. In many general acoustic networks, a lumped compliance
is used having a general impedance Z=1/iω·(V/ρma²). However the general method
in this thesis is more accurate and is also used for the connection A→B. Therefore,
the computation stations from A to E are distributed along the loop anti-clockwise,
which is shown in Fig. 4.3.3 (c). The computation orientation is shown in Fig. 4.3.3
(a), where the original point is located at the interface of the resonator tube and the
driver. Although in the tee section, acoustically generated flows and vortices may
occur, it is possible to neglect these in the acoustic limit, and only apply mass
conservation and equality of pressure in the tee. In that case, the joint tee is
simplified as a point.
Traveling-wave systems 93

p1out U1out
p1input p1 fb A
o
(a)
U1input U1 fb
(b) B
C
p1in D
E
U1in Resonator tube

o x (c)

Figure 4.3.3: Illustration of computation stations for a traveling-wave refrigerator


(a) the orientation of modeling computation: o indicates the origin (b) detailed
sketch of the joint tee (c) the global distribution of computation stations.

The computation stations are located at the interfaces between any two connected
components within the loop. O is the origin of the coordinate system, where the
driver is connected to the traveling-wave refrigerator part. The input acoustic flow
from the driver, feeding into the refrigerator at original point of O, is indicated
by p1in and U1in . The pressure and volume velocity at the end of the resonator tube
connecting the tee junction are indicated by p1input and U1input . The pressure and
volume velocity at one branch of the tee junction which is connected to the cold-
end heat exchanger are indicated by p1out and U1out . The pressure and volume
velocity at another branch of the tee junction which is connected to the feedback
inertance tube are indicated by p1 fb and U1 fb , see Fig. 4.3.3 (b) and (c). A locates
the interface of the feedback inertance tube and the compliance volume. B locates
the interface of the compliance volume and the ambient heat exchanger. C locates
the interface of the ambient heat exchanger and the regenerator. D locates the
interface of the regenerator and the cold end heat exchanger. E locates the interface
of the cold end heat exchanger and the tube section connecting the tee.
As a prerequisite, the pressure at the interface of the driver and the resonator tube
p1in and the cold end temperature TC of the regenerator or cold end heat exchanger
should be given. The cooling power and efficiency of the refrigerator system can
be computed by the model presented here.

Resonator tube
First, indicating the length of the resonator tube as Lres and the diameter as d res , it
is known that at x = 0 , by Eq. (4.3.19) and (4.3.20), we have
94 Chapter 4

p1in = C1res + C2 res , (4.3.21)


πd 2
U1in = − res
[C1res − C2res ] . (4.3.22)
4 aρ m
Thus the coefficients are
1 4U1in ρ m a 
C1res =  p1in −  , (4.3.23a)
2 πd res
2

1 4U1in ρ m a 
C2 res =  p1in +  . (4.3.23b)
2 πd res
2

Therefore, the pressure and volume velocity at x = Lres are
p1input = C1res ⋅ eikLres + C2 res ⋅ e- ikLres , (4.3.24)
πd res
[ ].
2
U1input = − C1res ⋅ eikL − C2 res ⋅ e- ikL
res res
(4.3.25)
4 aρ m
Substitution of Eq. (4.3.23a) and (4.3.23b) into (4.3.24) and (4.3.25), and splitting
the exponential factors into cosine and sine, yields
4U1in ρ m a
p1input = p1in cos(kLres ) − i sin(kLres ) , (4.3.26)
πd res
2

πd res
2
p1in
U1input = U1in cos(kLres ) − i sin(kLres ) . (4.3.27)
4ρma
Loop
For the loop, the computation starts from the right side of the joint tee point, where
the acoustic flow from the driver via resonator tube joins the acoustic flow out of
the cold-end heat exchanger, and merges into the feedback flow, indicated as
p1 fb and U1 fb . The computation goes stepwise through the inertance tube, A to E
till the tee branch connected to the cold-end heat exchanger. For the derivation
which is laborious, the reader is referred to appendix F. Finally the relation
between p1 fb U1 fb and p1out U1out is found:
 4 ρ aD 
p1out = p1 fb D1 cos(kLtb ) − i m 2 3 sin(kLtb )
 πd tb 
 4 ρ aD 
+ U1 fb D 2 cos(kLtb ) − i m 2 4 sin( kLtb ) , (4.3.28)
 πd tb 
 πd 2 D 
U1out = p1 fb D3 cos(kLtb ) − i tb 1 sin(kLtb )
 4ρma 
Traveling-wave systems 95

 πd 2 D 
+ U1 fb D4 cos(kLtb ) − i tb 2 sin(kLtb ) . (4.3.29)
 4ρma 
where the involved complex coefficients of D are
πd 2fb R0 f (τ , bµ )
D1 = (1 − iωC0 R0 g (τ , bµ ) )θ1 + i θ3 , (F.38a)
4ρma
4ρ a
D2 = − R0 f (τ , bµ )θ 2 − i (1 − iωC0 R0 g (τ , bµ ) ) m2 θ 4 , (F.38b)
πd fb
iωC0 ln τ iπd 2fb
D3 = θ1 − θ3 , (F.38c)
1−τ 4 ρ m aτ
4 ρ aωC ln τ
D4 = θ 2 / τ + m 2 0 θ , (F.38d)
πd fb (1 − τ ) 4
where temperature ratio τ (the ratio between the temperatures at both ends of the
regenerator) is defined by Eq.(F.18), and four real coefficients, which only depend
on the geometrical parameters, used in complex D s are:
θ1 = cos(kL fb ) cos(kLcpl ) − d 2fb sin(kL fb ) sin(kLcpl ) / d cpl
2
, (F.37a)
θ 2 = cos(kL fb ) cos(kLcpl ) − d cpl
2
sin(kL fb ) sin(kLcpl ) / d 2fb , (F.37b)
θ3 = sin(kL fb ) cos(kLcpl ) + d cos(kL fb ) sin(kLcpl ) / d
2
cpl
2
fb , (F.37c)
θ 4 = sin(kL fb ) cos(kLcpl ) + d cos(kL fb ) sin(kLcpl ) / d .
2
fb
2
cpl (F.37d)
Tee junction
The boundary conditions at the joint tee for the mode of thermoacoustic
refrigerator are
p1 fb = p1out = p1input , (4.3.30)
U1 fb = U1out + U1input . (4.3.31)
Applying boundary condition (4.3.30) to Eq. (4.3.28), it follows that
 4 ρ aD 
p1 fb 1 - D1 cos(kLtb ) + i m 2 3 sin(kLtb )
 πd tb 
 4 ρ aD 
= U1 fb D2 cos(kLtb ) − i m 2 4 sin(kLtb ) . (4.3.32)
 πd tb 
Thus, the acoustic impedance on the right hand side of the joint tee is given by:
4 ρ aD
D2 cos(kLtb ) − i m 2 4 sin(kLtb )
p1 fb πdtb
Z fb = = . (4.3.33)
U1 fb 1 - D cos(kL ) + i 4 ρ m aD3 sin(kL )
1 tb
πdtb2 tb

Applying boundary condition (4.3.31) to Eq. (4.3.29), it follows that


96 Chapter 4

 πd 2 D 
U1input = U1 fb − U1out = −p1 fb D3 cos(kLtb ) − i tb 1 sin(kLtb )
 4ρma 
 πd D
2

+ U1 fb 1 - D4 cos(kLtb ) + i tb 2 sin(kLtb ) . (4.3.34)
 4ρma 
By using Eq. (4.3.30) and (4.3.34), the acoustic impedance of the joint tee is given
by
p1input p1 fb
Z input = = =
U1input U1 fb − U1out
1
.
1  πd D2 2
 πdtb2 D1
1 - D4 cos(kLtb ) + i sin(kLtb ) − D3 cos(kLtb ) + i
tb
sin(kLtb )
Z fb  4ρma  4ρma
(4.3.35)
Substitution of Eq. (4.3.33) into (4.3.35) yields
4 ρ aD
D2 cos(kLtb ) − i m 2 4 sin(kLtb )
πdtb
Zinput = .
 πdtb2 D2 4 ρ m aD3 
1 + D1D4 − D2 D3 − (D1 + D4 ) cos(kLtb ) + i +  sin(kLtb )
 4ρma πdtb2 
(4.3.36)
So this is the input impedance of the loop seen from the connection which connects
the resonator tube and the tee junction.
The complete system
Using the transmission relation for the acoustic impedance, see appendix G,
replacing Z c with Z input in Eq. (G.10) yields the complete acoustic impedance of
the refrigerator
4ρma
Z input cos(kLres ) + i sin(kLres )
4ρma πd res
2
Z a − rfg = . (4.3.37)
πd res
2
4ρma
cos(kLres ) + iZ input sin(kLres )
πd res
2

It is noticeable that the complete acoustic impedance of the refrigerator Z a − rfg only
depends on the geometrical configuration of the refrigerator and operational
temperatures.
Assume that the pressure at the interface between the driver and the resonator tube
is known, i.e. p1in from the driver being known. Thus, the volume velocity U1in is
given by
U1in = p1in / Z a − rfg . (4.3.38)
Traveling-wave systems 97

Substitution p1in and U1in into Eq. (4.3.26) and (4.3.27) yields
4p1in ρ m a
p1input = p1in cos(kLres ) − i sin(kLres ) , (4.3.39)
πd res
2
Z a − rfg
πd res
2
p1in
U1input = p1in cos(kLres ) / Z a−rfg − i sin(kLres ) . (4.3.40)
4ρma
By using the boundary condition (4.3.30) and Eq. (4.3.39), the pressure p1 fb is
given by
 4ρ a 
p1 fb = p1in cos(kLres ) − i 2 m sin(kLres ) . (4.3.41)
 πd res Z a−rfg 
By using Eq. (4.3.33), the volume velocity U1 fb is given by

p1in  4ρma 
U1 fb = cos(kLres ) − i 2 sin(kLres ) . (4.3.42)
Z fb  πd res Z a − rfg 
Now, all equations derived have to be reviewed to understand the result of the full
analysis. If the output pressure at the driver p1in which forms the input to the
system is given, the key reference parameters p1 fb and U1 fb will be given by Eq.
(4.3.41) and (4.3.42). From this all the other acoustic pressures and volume
velocities at all the computation stations from A to E can be obtained by the
corresponding equations in appendix F.
Energy flows
Since the acoustic field is now known, the cooling power can be computed. As
mentioned in the beginning of this section, the cooling power is given by Eq. (4.3.4)
Q& C = W& dn − st − E& 2 .
It can be seen from Fig.4.3.3, and using Eq. (D.2), the cooling power can be written
as
1 ~
Q& C = Re[p1D U1D ] − E& 2 . (4.3.43)
2
Substitution of Eq. (F.34) and (F.35) into the first term of Eq. (4.3.43) on the right
hand side yields

2
~ ~ ~ 
1 ~ 1 ~ Re[D2 D4 ] D2 D3 D1D4 
Re[p1D U1D ] = p1 fb  Re[D1D3 ] + 2
+ Re[ + ~ ] .
2 2  Z Z fb Z fb 
 fb 
(4.3.44)
The term between brackets is defined as Θ3
98 Chapter 4

~ ~ ~
~ Re[D2 D4 ] D2 D3 D1D4
Θ3 = Re[D1D3 ] + 2
+ Re[ + ~ ]. (4.3.45)
Z fb Z fb Z fb
Thus, Eq. (4.3.44) can be written as
1 ~ 1 2
Re[p1D U1D ] = p1 fb ⋅ Θ3 . (4.3.46)
2 2
Using the distributed regenerator model Eq. (F.16) at position C, and noticing Eq.
(F.9) and Tm positionC = T0 , the pressure gradient at position C is given by
dp1 R0
=− U1B ( x) . (4.3.47)
dx positionC Lreg
By using Eq. (2.1.72) and (F.17), and assuming that the regenerator part is ideally
thermally isolated from the ambient environment and the heat capacity of the solid
material of the regenerator is large enough, therefore the constant total energy flow
along the regenerator can be evaluated at position C

E2 = −
& ψ reg Areg R0  ~  ~
Im p1B U1B 1 − fν −
(fκ − fν
~
) 

2ωρ m Lreg   (1 + ε s )(1 + σ ) 

+
ψ reg Areg c pT0 R02 1 / τ − 1 ~
U1B Im  fν +
2 ( )
fκ − fν (1 + ε s fν / fκ ) 
~

2ω 3 ρ m (1 − σ ) L2reg Lreg  (1 + ε s )(1 + σ ) 


1/ τ − 1
[ ]
− ψ reg K + (1.0 − ψ reg ) K s AregT0
Lreg
. (4.3.48)

Note that for an ideal regenerator, so with mesh size much below the thermal
penetration depth of the acoustic wave, and without heat conduction, it follows
that E& 2 = 0 , as it should be.
Substitution of Eq. (4.3.46) and (4.3.48) into Eq. (4.3.43) yields
1 − (1 / τ )
1
2
2
[ ]
Q& C = p1 fb ⋅ Θ3 − ψ reg K + (1.0 − ψ reg ) K s ⋅ AregT0
Lreg
ψ A R  ~ 
+ reg reg 0 Im p1B U1B 1 − fν −
~ (
fκ − fν
~
) 

2ωρ m Lreg   (1 + ε s )(1 + σ ) 



ψ reg Areg c pT0 R02 1 / τ − 1 ~
U1B Im  fν +
2 ( )
fκ − fν (1 + ε s fν / fκ ) 
~

(1 + ε s )(1 + σ ) 
.
2ω 3 ρ m (1 − σ ) L2reg Lreg 
(4.3.49)
Substitution of Eq. (F.28) and (F.29) into Eq. (4.3.49) makes Q& C be expressed by
p1 fb and Z fb (eliminate p1B and U1B ). The cooling power is given in form of p1 fb
and Z fb
Traveling-wave systems 99

1 − (1 / τ )
1
2
2
[
Q& C = p1 fb ⋅ Θ3 − ψ reg K + (1.0 − ψ reg ) K s AregT0 ⋅] Lreg

+
ψ reg Areg R0 2  θ θ θ θ 

p1 fb Im  ~ + 3 4  ~ Tm β fκ − fν 
~

( )
1 − fν − (1 + ε )(1 + σ ) 
1 2
2ωρ m Lreg  
 fb
Z Z fb  s 

+
ψ reg Areg R0  2
2  πd fbθ1θ 3

4 ρ aθ θ   ~ T β f − fν 
+ m 2 24  Re 1 − fν − m κ
( ~
)
(1 + ε s )(1 + σ )
p1 fb 
2ωρ m Lreg  4ρma πd 2fb Z fb  

ψ reg Areg c pT0 R02 1 / τ − 1  2
2 θ  πd 2fbθ 3  πd 2fbθ 2θ3 Im[Z fb ] 
2

− 3 p1 fb  2
+   +
2ω ρ m (1 − σ ) L2reg Lreg  Z fb
2
 4 ρ a 
 2ρm a 2 
Z fb 
 m

~
⋅ Im  fν + κ
( ~
)
f − fν (1 + ε s fν / fκ ) 
(1 + ε s )(1 + σ ) 
. (4.3.50)

Using Eq. (4.3.41), it is obtained
  4 ρ a sin(kL ) 
2

p1 fb = p1in (cos(kLres ) ) +  m 2 res 
2 2 2

  πd Z 
 res a − rfg 
4 ρ a sin(2kLres )  1  
+ m Im   . (4.3.51)
πd res
2
 Z a − rfg  
Substitution of Eq. (4.3.51) into (4.3.50), the cooling power can be rewritten as
1 − (1 / τ )
2
[
Q& C = p1in Θ1{Θ3 / 2 + Θ2 } − ψ reg K + (1.0 − ψ reg ) K s AregT0 ]
Lreg
,

(4.3.52)
where the new functions Θ1 and Θ 2 are defined as:
2
 4 ρ a sin(kL )  4 ρ a sin(2kL )  1 
Θ1 = (cos(kLres ) ) +  m 2 res 
+ m
2 res
Im  
 πd res Z a − rfg 
 
π 2
d res  Z a − rfg 
(4.3.53)

Θ2 =
ψ reg Areg R0  θ θ θ θ 

Im  ~ + 1 2 3 4  ~ Tm β fκ − fν 
1 − fν −
( ~

)
2ωρ m Lreg  Z fb Z fb  (1 + ε s )(1 + σ ) 

+
ψ reg Areg R0  πd 2fbθ1θ3 4 ρ m aθ 2θ 4   ~ Tm β fκ − fν 
+ Re 1 − fν −
~
( )
2ωρ m Lreg  4 ρ m a πd 2
Z
2 
  (1 + ε s )(1 + σ )
 fb fb 
100 Chapter 4

ψ reg Areg c pT0 R02 1 / τ − 1  θ 22  πd 2fbθ 3  πd 2fbθ 2θ3 Im[Z fb ] 


2

− 3 +  +
2ω ρ m (1 − σ ) L2reg Lreg  Z 2  4 ρ m a  2ρm a 2 
Z fb 
 fb 
~ (
⋅ Im  fν + κ
~
)
f − fν (1 + ε s fν / fκ ) 
(1 + ε s )(1 + σ ) 
. (4.3.54)

The functions Θ1 , Θ 2 and Θ3 only depend on the specific configuration and working
temperature conditions. It should be emphasized that this final result is obtained
under some assumptions, which are stated during derivation: no blockage ideal heat
exchangers; no losses outward; no mean fluid flow; no temperature gradient along
the pipes except the regenerator.
Coefficient of performance (COP)
From Eqs. (4.3.26) and (4.3.27), it can be obtained that the input acoustic work to
the system from the driver is
1
2
[
~ 1
2
] [~
W&in = Re p1in U1in = Re p1input U1input . ] (4.3.55)

Using Eqs. (4.3.33) and (4.3.34), the volume velocity can be rewritten as
 πd tb2 D1 
U1input = U1 fb − U1out = −p1 fb D3 cos(kLtb ) − i sin(kLtb )
 4ρma 
p  πd D2

+ 1 fb 1 - D 4 cos(kLtb ) + i tb 2 sin(kLtb ) . (4.3.56)
Z fb  4ρm a 
Substitution of Eq. (4.3.30), (4.3.33) and (4.3.56) into (4.3.55) yields
1 2
W&in = p1 fb ⋅ Θ 4 , (4.3.57)
2
where Θ 4 is defined as
 ~
 4 ρ m aD 3  ~
Θ 4 = Re (1 - D4 cos(kLtb ) )1 − i sin(kLtb )  + D1 (D4 - cos(kLtb ) )
  πdtb 2

~
πd 2 D  ~ 4 ρ aD 
− i tb 2 sin( kLtb ) /  D2 cos(kLtb ) + i m 2 4 sin(kLtb )  . (4.3.58)
4ρma   πdtb 
Using Eq. (4.3.51), the input acoustic power can be rewritten as
1
W&in = p1in ⋅ Θ1 ⋅ Θ 4 .
2
(4.3.59)
2
Thus, by the definition of coefficient of performance for a refrigerator, it is given
by

COP = = −
[ ]
Q& C Θ3 + 2Θ 2 2 ψ reg K + (1.0 − ψ reg ) K s AregT0 1 − (1 / τ )
. (4.3.60)
W&in Θ4 p1in ⋅ Θ1Θ 4
2
Lreg
Traveling-wave systems 101

Despite the laborious amount of computation these expressions give an analytical


means of studying a traveling-wave system. So if the configuration and the
temperatures at both ends of the regenerator are known, the efficiencies and powers
are readily computed.
Computational validation of the model
To test the analytical model, a comparison between results, computed by the
proposed analytical equations, and by DeltaE [38] is made for the thermoacoustic
engine described in S. Backhaus and G. W. Swift’s paper in 2000 [34] and
described in the appendix of his book [38], as shown in Fig. 4.3.6.
The system consists of a ¼-wavelength resonator filled with 30-bar helium, a torus-
shaped section containing a regenerator that is sandwiched by two heat exchangers,
and a variable acoustic load. It will be described in detail below.
The main components are as listed:
1→2: The main cold heat exchanger is of shell-and-tube construction consisting of
299 stainless-steel tubes with 2.5-mm inside-diameter with wall thickness of 0.7
mm, and length of 20-mm-long tubes welded into two 1.6-mm-thick stainless-steel
plates. The diameter of the heat exchanger is 9.5 cm. It is cooled by chilled water
of 15°C.
2→3: The regenerator is made from a 7.3-cm-tall stack of stainless-steel wire
screens machined to a diameter of 8.89 cm. The wire screen has mesh 120 and wire
diameter of 65 µm. These wire screens are contained by a thin-wall stainless-steel
can. The detailed description about this kind of regenerator material, stainless wire
screen, and its relative parameters will be explained later in section 4.4.2.2. The
volume porosity of the regenerator is given as 0.72 and the hydraulic radius is
given as 42 µm in the paper [34].
3→4: The hot heat exchanger consists of a 0.64-cm-wide by 3.5-m-long Ni-Cr
ribbon wound zigzag on an alumina frame. The ribbon is divided into 3 equal-
length segments and driven with 208-V three-phase power in a delta configuration.
4→5: The thermal buffer tube (TBT) is a tapered, 24-cm-long open cylinder tube
of Inconel 625. The wall thickness is 4.0 mm. The upper 8.0 cm of the TBT is a
straight cylinder while the lower 16.0 cm is flared in diameter from 8.9 cm to 9.6
cm with a 1.35° half-angle taper.
5→6: The secondary heat exchanger is water-cooled and of shell-and-tube, used to
maintain the lower end of TBT at room temperature. It contains 109 4.6-mm
inside-diameter, 10-mm-long stainless-steel tubes welded into two 1.6-mm-thick
stainless-steel plates.
6→7→8: The junction is a standard-wall, 3 ½-in. nominal, stainless-steel tee. The
inside diameter is around 9.0 cm.
102 Chapter 4

11 12
10 1
2
3
4
5
9 6

8 7 13 14 15 16

Figure 4.3.6: A schematic drawing of the engine, resonator and variable acoustic
load. a and b are figures from reference paper [34]. c is schetch of the whole
engine.
Traveling-wave systems 103

8→11: The feedback inertance is composed of three separate sections. The first
section (8→9) is a 3 ½-in. to 3-in. nominal. long-radius reducing elbow. The
centerline length of the elbow is 20.9 cm, and the final inside diameter of the elbow
is 7.8 cm. The second section (9→10) is a 3-in. nominal, stainless-steel pipe of
25.6 cm long. The third section (9→10) is a machined cone that adapts the 3-in.
nominal pipe to the compliance. Its initial inside diameter is 7.8 cm and end is
enlarged to 10.2 cm by a taper angle of 13.5°. Its length is 10.2 cm.
11→12: The compliance consists of two 4-in. nominal, short-radius 90° elbows
made from carbon steel. Its internal volume is 2830 cm³.
12→1: Between the compliance and the main cold heat exchanger is a device
termed as “jet pump”, used to prevent Gedeon streaming.
7→16: The resonator consists of three sections. The first section (7→13) is a
machined cone that adapts the 3 ½-in. nominal tee to a 4-in. nominal, carbon-steel
pipe. The initial inside diameter 9.0 cm is enlarged to 10.2 cm over a length of 10.2
cm, giving a 6.8° taper angle. The main section of the resonator (13→14) is a 1.9-
m length of 4-in. nominal, carbon-steel pipe. The inside diameter is 10.2 cm. The
last section of the resonator includes a 7° cone (14→15) which enlarges the inside
diameter of the resonator from 10.2 cm to 25.5 cm over a length of 1.22 m. The
large end of the cone is closed with a 25.5-cm-diameter pipe with an approximate
length of 52 cm, terminated by a 2:1 ellipsoidal cap (15→16).

The operation condition described in Swift’s book [38], page 272-280 of appendix
B.4, is used as comparison example. As the acoustic load is not clear for the author
in the computation case in the book [38], the acoustic pressure at the junction tee is
designated as the input parameter p1in for the computation, which is given as
p = 2.7231E + 05 Pa in the book [38]. The temperatures at the main cold heat
exchanger and hot heat exchanger are given as ambient temperature 300 K and 900
K, respectively. The working frequency and mean pressure are given in the book:
f=84.12 Hz and Pm=3103 kPa. The necessary geometrical parameters are used as
listed in the above paragraph. Gas model and properties like viscosity is described
in the model in section 4.3.
The code is built in Fortran (appendix H) and follows the sequence:
1. Input of the geometrical parameters, ambient temperature 300 K and hot heat
exchanger temperature 900 K, using Eq. (F.38 a b c d) and (F.39 a b c d) to
compute the real and complex coefficients θ1 θ2 θ3 θ4 D1 D2 D3 D4 .
2. Use Eq. (4.3.33) and (4.3.36) to obtain Zfb and Zinput .
3. The volume velocities are given by
U1input = p1in / Z input and U1 fb = p1in / Z fb .
4. Use Eq. (4.3.28) and (4.3.29) to compute P1out and U1out .
104 Chapter 4

The results using the analytical lumped-element model discribed by Backhaus and
Swift [34] are also presented in the following table as comparison.

Results from the Results from


related analytical lumped-element
DeltaE
equations of this model [ 34]
chapter
Heating power
into hot heat 3037 3108 11042
exchanger [W]
U1input
(amplitude,
0.29, -89.5 0.26, -83.7 0.17, -62.2
phase) [m³/s,
degree]
U1 fb
(amplitude,
0.20, 86.4 0.18, 84.1 0.08, 69.3
phase) [m³/s,
degree]
U1out
(amplitude,
8.63E-2, -77.6 9.21E-2, -59.9 0.96E-2, -20.7
phase) [m³/s,
degree]
U1location 2
(amplitude,
8.56E-3, 12.2 8.65E-3, 12.7 2.79E-2, -19.9
phase) [m³/s,
degree]
U1location 3
(amplitude,
3.02E-2, -46.6 3.02E-2, -46.4 0.96E-2, -20.7
phase) [m³/s,
degree]

Table 4.3.I: Comparison of computation from DeltaE, analytical model developed


in this chapter, and lumped-element model [34].

The volume velocities at the three branches of the tee, shown in figure 4.3.7, and
the input heating power at the hot heat exchanger are computed by the analytical
model described in the previous section and also by the lumped-element in
reference [34]. The flows at location 2 and 3 are given and compared as well.
Traveling-wave systems 105

p1out U1out
p1input p1 fb
U1input U1 fb
p1in

Figure 4.3.7: A schematic drawing of the tee and pressures and velocities in three
branches.

The results are listed in table 4.3.I. The results from the proposed analytical model
and those from DeltaE provided by the book [38] agree well with each other. But
the results from the lumped-element model differ much from those of DeltaE.
Therefore, the proposed analytical model is more accurate than the lumped-element
model.

Prediction of influence of geometrical parameters on the performance of an


example refrigerator
The thermoacoustic engine of Backhaus and Swift’s [34], as shown in Fig. 4.3.6,
was assumed to operate in the cooling mode, used as an example refrigerator in this
computation. The geometry can be found in the description following Fig. 4.3.6. In
the refrigerator mode, the main cold heat exchanger (1→2) works as an ambient
heat exchanger and the hot heat exchanger (3→4) becomes cold heat exchanger.
The resonator tube (7→16) was a quarter-wave-length tube is Swift’s engine. In
this computation, the resonator tube (7→16) was extended to a half-wave-length
tube. It is assumed that input acoustic pressure p1in at one end of the resonator
connecting the driver is constant p = 2.7231E + 05 Pa. The working frequency and
mean pressure for the engine mode given in the book [38]: f=84.12 Hz and
Pm=3103 kPa remain the same. The working gas is helium like in the engine mode.
Assume that this refrigerator is driven by a driver and the ambient heat exchanger
is maintained at environmental temperature around 300 K. From Eq. (4.3.37), it
can be concluded that the complete acoustic impedance of the refrigerator only
depends on the geometry and working condition parameters (TC, Pm and working
gas). If the cold-end temperature is given, the influence of any of the relevant
geometrical parameters on the impedance can be investigated by using Eq. (4.3.37).
We assume that the cold-end temperature is 200 K. For every computation, only
one geometrical parameter is varied at constant further settings. The geometrical
parameters used in Eq. (4.3.37) (see appendix F) which are varied individually at
each time in computation are listed in the table 4.3.II. In table 4.3.II, the parameters
having a big influence on the total acoustic impedance are marked by “(!!)”; less
important parameters for impedance are marked with “(!)”. Those parameters
106 Chapter 4

having negligible influence are not marked. The important parameters are
discussed below as well as the operating frequency.

hydraulic
component length diameter porosity
radius rh
resonator tube Lres (!!) dres × ×
feedback inertance tube Lfb (!!) dfb (!) × ×
compliance tube Lcpl (!) dcpl (!) × ×
regenerator Lreg (!!) dreg (!) rh-reg (!!) ψreg (!!)
tube segment from the cold
end heat exchanger to the Ltb dtb × ×
center of tee junction
Table 4.3.II: Geometrical parameters in Eq. (4.3.37) and appendix F.

1. Dependency on the resonator tube length


If only the resonator tube length is varied, the complete acoustic impedance
changes in amplitude and phase as shown in Figs. 4.3.8 and 4.3.9. The effect on
COP and relative COP is given in Fig. 4.3.10.
As seen in Figs. 4.3.8 and 4.3.9, both the amplitude and phase of the complete
acoustic impedance have a large variation with the resonator tube length. So tuning
of this length in experimental set-ups will be important.
Log(amplitude of complete acoustic impedance

7.5

7.0
of the refrigerator Za_rfg)

6.5

6.0

5.5

5.0

4.5

4.0
0 1 2 3 4 5 6 7 8

Resonator tube length (m)

Figure 4.3.8: The amplitude of the complete acoustic impedance of the example
refrigerator as a function of the resonator tube length.
Traveling-wave systems 107

100

Phase of complete acoustic impedance


80

60

of the refrigerator Za-rfg


40

20

0
0 1 2 3 4 5 6 7 8
-20
Resonator tube length (m)
-40

-60

-80

-100

Figure 4.3.9: The phase of the complete acoustic impedance of the example
refrigerator as a function of the resonator tube length.

0.85

0.80

0.75

0.70

0.65
COP
COP

0.60 Relative COP to Carnot's


0.55

0.50

0.45

0.40

0 1 2 3 4 5 6 7 8
Resonator tube length (m)

Figure 4.3.10: COP and relative COP as a function of the resonator tube length.

In Fig.4.3.10, the COP and relative COP do not show a strong dependency on the
resonator tube length. With respect to the acoustic coupling between driver and
thermoacoustic device, it is vital to choose the suitable resonator length in the
design.
108 Chapter 4

2. Dependency on the feedback inertance tube


Now the feedback inertance tube length is varied. The original geometrical
parameters except the length of the feedback inertance tube in Eq. (4.3.37) and
related other equations are used. The corresponding complete acoustic impedances
at various lengths of the feedback inertance tube are obtained. Although the
variation of the amplitude of the complete acoustic impedance is not so large in Fig.
4.3.8, the phase shows a strong dependency. The variation of COP and relative
COP is given in Fig. 4.3.11. As seen in Fig. 4.3.11, the COP and relative COP have
strong dependency on the feedback tube length. Although Fig. 4.3.11 shows a high
COP at zero feedback inertance tube length, it leads to a zero cooling power and a
zero input acoustic power in practical operation. This effect is also shown by the
computation results given by Fig. 4.3.12. Therefore, zero-feedback-tube-length is
not a realistic option for the design of a traveling-wave refrigerator.

2
COP
Relative COP to Carnot's
COP

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Feedback inertance tube length (m)
-1

-2

-3

-4

Figure 4.3.11: COP and relative COP as a function of the feedback inertance tube
length.

If only the diameter of the feedback inertance tube is varied, the results show that
the complete acoustic impedance has a small variation of both the amplitude and
the phase (about 4 degrees difference if the diameter varies from 0.04 m to 0.16 m).
Traveling-wave systems 109

50000

40000

30000 cooling power


input acoustic power
20000

Power (W)
10000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
-10000 Feedback inertance tube length (m)
-20000

-30000

-40000

-50000

Figure 4.3.12: Cooling power and input acoustic power as a function of the
feedback inertance tube length.

3. Dependency on the compliance tube


Now the length and the diameter of the compliance tube are varied each at one time,
the results show that the complete acoustic impedance has small variations in the
amplitude and variation in the order of around 10-15 degrees in the phase on both
the length (vary in the range of 0.1 m to 1.2 m) and the diameter of the compliance
tube (vary in the range of 0.04 m to 0.2 m).

4. Dependency on the regenerator


The computation has shown that the complete acoustic impedance is slightly
dependent on the length, diameter and porosity of the regenerator, and the
hydraulic radius of the wire gauzes. The results show small variation in the
amplitude and the phase (around 4-5 degrees difference). That might be attributed
to the small variation ranges for these parameters in computation by considering
the practice. The regenerator is still a key component and important for further
investigation. The influence of two important parameters of the regenerator
material, porosity and hydraulic radius, on the COP and relative COP are given in
Figs. 4.3.13 and 4.3.14. In Fig. 4.3.13, the efficiency increases with increasing
porosity. In Fig. 4.3.14, the curve shows that there exists an optimum value of
hydraulic radius, which makes a maximum efficiency. Fig. 4.3.14 shows a flat peak
over a broad range of the hydraulic radius. These two important parameters will be
discussed more in detail theoretically and experimentally in the following section
4.4.
110 Chapter 4

0.9

0.8

0.7
COP

COP
Relative COP to Carnot's
0.6

0.5

0.4

0.4 0.5 0.6 0.7 0.8 0.9 1.0


Porosity of regenerator

Figure 4.3.13: COP and relative COP as a function of the porosity of the
regenerator.

2.0

1.5
COP

1.0

0.5

0.0
0 10 20 30 40 50 60
-0.5 Hydraulic radius of regenerator (µm)

-1.0
COP
-1.5
Relative COP to Carnot's

-2.0

Figure 4.3.14: COP and relative COP as a function of the hydraulic radius of the
regenerator material.

The influence of the regenerator length on the COP and relative COP are given in
Figs. 4.3.15. The influence on cooling power and input acoustic power is given in
Fig. 4.3.16. Fig. 4.3.15 shows that the regenerator has to be longer than a certain
value to gain positive COP, and the COP increases with increasing regenerator
length. When the regenerator length is larger than a certain value, the COPs almost
Traveling-wave systems 111

stay constant. In Fig. 4.3.16, the cooling power and input acoustic power curves
first increase with increasing regenerator length, then decrease slowly.

1
COP

0
0.00 0.05 0.10 0.15 0.20 0.25
Regenerator length (m)

-1
COP
Relative COP to Carnot's

-2

Figure 4.3.15: COP and relative COP as a function of the regenerator length.

5000

4000
Power (W)

3000

2000
cooling power
input acoustic power
1000

0
0.00 0.05 0.10 0.15 0.20 0.25
Regenerator length (m)
-1000

Figure 4.3.16: Cooling power and input acoustic power as a function of the
regenerator length.

5. Dependency on the tube segment from the cold end heat exchanger to the center
of tee junction
The results have shown that the complete acoustic impedance is slightly dependent
on the length and diameter of this tube segment.
112 Chapter 4

6. Dependency on operating frequency


The original geometrical parameters were used and the cold end temperature was
set as 200 K. Helium gas at filling pressure of Pm=3103 kPa is used. The operating
frequencies were varied as input into Eq. (4.3.37) and the complete acoustic
impedance changed in amplitude and phase as shown in Fig. 4.3.17 and 4.3.18.
In Figs. 4.3.17 and 4.3.18, the amplitude and phase of the complete acoustic
impedance have a large variation with operating frequency, which are analogous to
Figs.4.3.8 and 4.3.9. This shows that there is strong coupling between the resonator
tube length and operating frequency. The influence of operating frequency on the
efficiency and relative efficiency is given in Fig.4.3.19. Fig.4.3.19 shows that an
optimum operating frequency exists.
Log(amplitude of complete acoustic impedance

7.5

7.0
of the refrigerator Za_rfg)

6.5

6.0

5.5

5.0

4.5

4.0

3.5
0 20 40 60 80 100 120 140 160 180
Frequency (Hz)

Figure 4.3.17: The amplitude of the complete acoustic impedance of the example
refrigerator as a function of operating frequency.
Traveling-wave systems 113

100

Phase of complete acoustic impedance


80

of the refrigerator Za-rfg


60

40

20

0
0 20 40 60 80 100 120 140 160 180
-20 Frequency (Hz)
-40

-60

-80

-100

Figure 4.3.18: The phase of the complete acoustic impedance of the example
refrigerator as a function of operating frequency.

3
COP
Relative COP to Carnot's
2
COP

0
0 20 40 60 80 100 120 140 160 180
Frequency (Hz)
-1

-2

-3

Figure 4.3.19: COP and relative COP as a function of operating frequency.


114 Chapter 4

4.4 Optimizing regenerator material

4.4.1 Introduction
The regenerator is a key component in a traveling-wave system, where the heat
exchange between the working gas and the solid filling material takes place to
realize the thermal cycle. A regenerator consists of a porous medium. A variety of
porous materials can be chosen: ceramic honeycomb with different channel shapes,
metal honeycomb, stainless steel wire screens, for instance. Therefore, the thermal
and geometrical properties of the regenerator materials are of great interest in
optimizing the performance of a traveling-wave system. In this section, the
optimization of regenerator material is investigated experimentally. Samples of
ceramic and metal honeycomb in different channel sizes, and stainless steel wire
screens in different hydraulic radii and porosities, were applied in a coaxial
traveling-wave engine to characterize their performance. In practice, stainless steel
wire screens are widely used as filling material for regenerators. Therefore,
stainless steel screens are the main concern of this experiment.
This work is an extension of the TASOR project (Thermo Acoustic Systems to
Upgrade Waste heat) which is supported by Senter Novem. The TASOR project is
cooperation with the Energy research Center of the Netherlands (ECN) who
coordinated the project, Huisman Elektrotechniek, Aster Thermoacoustics, NRG
who makes the computational fluid dynamic calculations and the Eindhoven
University of Technology (TUE) who is responsible for the research to
regenerators of a thermoacoustic engine.

4.4.2 Regenerator materials


4.4.2.1 Ceramic and metal honeycomb regenerator
Ceramic and metal honeycomb regenerator materials can differ widely in the
shapes and sizes of channels. Photos of two samples of ceramic and metal
honeycomb regenerators are given in Fig. 4.4.1. In this study, the tested
honeycomb samples all have the same geometry. They are cylinders consisting of
many parallel channels with square-shaped cross-section, as sketched in Fig.4.4.2.
Based on periodicity the geometry can be reduced to one single cell for analysis.
The length of the cells is indicated as Lreg . For one cell, the inner size is denoted
as d , and the wall-thickness is denoted as δ .

CPSI
CPSI, “cells per square inch” in full, is an important quantity to characterize the
amount of cells per unit area of cross section. For square shaped cells, it can be
calculated as
Traveling-wave systems 115

2
 25.4 
CPSI =   , (4.4.1)
d +δ 
where d and δ are specified in mm.

Figure 4.4.1: Photos of a ceramic and a metal honeycomb regenerators.

Lreg
One cell

Figure 4.4.2: Sketch of a honeycomb regenerator.

Hydraulic radius and hydraulic diameter


According to the definition of hydraulic radius for flow through a channel with
arbitrary geometry, which is the ratio between the volume occupied by fluid to the
“wetted area”, the hydraulic radius of square shape channel honeycomb regenerator
is evaluated as
V fluid d 2 Lreg d
rh = = = . (4.4.2)
Awetted 4 ⋅ d ⋅ Lreg 4
The hydraulic diameter is equal to four times the hydraulic radius:
116 Chapter 4

V fluid
Dh = 4 =d. (4.4.3)
Awetted

Porosity
The porosity is defined as the ratio between the volume, occupied by fluid, to the
total volume of the regenerator.
It is sufficient to consider only one cell. Thus, the porosity can be expressed as
V fluid d 2 ⋅ Lreg d2
ψ reg = = = .
Vtotal (d + δ )2 ⋅ Lreg (d + δ )2
(4.4.4)

Sometimes this is also named as OFA, open fluid area.

4.4.2.2 Regenerator of stainless steel wire screens


For wire screen regenerators, the situation is more complex than for square shaped
channel honeycombs. The wires are curved, and one sheet overlaps another one in
a random way when they are packed into a regenerator. A photo of regenerator
screens and the top view and side view of one screen are given in Fig.4.4.3.
Some assumptions have to be made before calculating the parameters hydraulic
radius and porosity. The diameter of the wires is indicated as δ and the inner size
of one cell is noted as d . A sketch of one cell is given in Fig.4.4.4 (a). We assume
that the wires of a screen are not curved, but perfect cylinders. Thus, a cell is
reduced to a structure confined by four half-cylinders (the wires) perpendicular to
each other at the corners. The length of one single half-cylinder is d + δ , as shown
in Fig.4.4.4 (b). A cell in principle has a thickness of twice the wire diameter
because of the crossing of the wires. In order to find a realistic estimation of one
cell thickness, the pattern how several screens are packed has to be investigated.
For randomly packed screens, they can be shifted or rotated with respect to each
other, so that the squared cells do not lie just above each other, as shown in Fig.
4.4.5 (a). In the following, it is assumed that the squared spacings can be shifted in
only two directions, shown in Fig. 4.4.5 (b), not allowed to rotate.
This shifting means that the square spacings can partially be seated into each other,
if more screens are stacked up. Thus, the effective thickness of one screen layer is
smaller than 2δ , as shown in the lower line of Fig. 4.4.5. The effective thickness is
evaluated by γ pf , which is defined as the screen packing density. So γ pf = 2
corresponds with the non-shifted situation. The normal value of γ pf is expected to
be around 2.
By using the above modeling for wire screen regenerators, the hydraulic radius,
hydraulic diameter and porosity can be obtainable as shown below.
Traveling-wave systems 117

(b)

(a) (c)

Figure 4.4.3: (a) Photo of regenerator screens and its top (b) and side views (c) of
one screen.

(a) One cell and its (b) One cell under assumptions and its
characteristic lengths characteristic lengths
Figure 4.4.4: A sketch of one cell in a wire screen regenerator and the simplified
unit cell.

Mesh
Similar to CPSI used in honeycomb regenerators, mesh, M is frequently used to
count the number of aligned cells per inch in a screen. It is obtained by
25.4
M= , (4.4.5)
d +δ
where d and δ are in mm.
118 Chapter 4

γ pf δ γ pf δ

Figure 4.4.5: Sketches of the packing of the wire screens. The upper line describes
the case where the screens are perfectly aligned so that every screen layer
occupies twice the wire thickness. The lower line describes the case where the
screens are shifted in only two directions so that one screen layer occupies less
than twice the wire thickness.

Hydraulic radius and diameter


The fact that the cells are not aligned with each other due to screens shifting is not
taken into account, but the effective thickness regarding to the factor γ caused by
shifting is considered. By this assumption, the complex structure is reduced to the
repetition of one cell so that one cell is sufficient to be considered instead of the
whole regenerator.
Then, the total volume occupied by one cell is
Vtotal = γ pf δ ⋅ (d + δ ) .
2
(4.4.6)
The volume of gas equals the total volume minus the volume of wires:
1 πδ 2
V fluid = Vtotal − Vwires = γ pf δ ⋅ (d + δ ) − 4⋅ ⋅ (d + δ ) .
2
(4.4.7)
2 4
The wetted area is
Awetted = 4 ⋅ π δ (d + δ ) .
1
(4.4.8)
2
By using the definition of hydraulic radius and diameter, they are expressed as:
Traveling-wave systems 119

V fluid γ (d + δ ) − π δ / 2
rh = = pf , (4.4.9)
Awetted 2π
V 4γ (d + δ ) − 2π δ
Dh = 4 fluid = pf . (4.4.10)
Awetted 2π

Porosity
By similar calculations, the porosity is given by
V fluid πδ
ψ reg = =1− .
2γ pf (d + δ )
(4.4.11)
Vtotal

4.4.2.3 Thermal conductivity


By using Fourier’s law, the heat flux in the axial direction is
dTm
q& = − K eff , (4.4.12)
dx
where K eff is effective thermal conductivity of the porous regenerator.
If the regenerator has a cross sectional area A , as shown in Fig.4.4.6 (a), the total
heat flow Q& cond through the regenerator is
dT
Q& cond = − A K eff m . (4.4.13)
dx

Lreg

Solid

A
Fluid

TCold THot (b) The regenerator modeled as two


(a) A regeneraor parallel electrical resistors.

Figure 4.4.6: Sketches of a regenerator and the analogous model as two parallel
electrical resistors.

In practical situation, the heat flow is complicated. Because the solid part is not a
massive rod and the fluid is not homogeneous tube flow. In this work, the heat flow
is modeled analogous to a current in a resistor. The fluxes through the thermal
120 Chapter 4

resistance of solid part and working gas are analogous to current in parallel
resistors, as shown in Fig. 4.4.6 (b). Assume that the solid part and the working gas
have the same temperature gradient in the x direction. Thus, the total heat flow
consists of heat flow through the solid part and working gas:
dT dT
Q& cond = Q& solid + Q& fluid = − Asolid K s m − A fluid K m . (4.4.14)
dx dx
By using the porosity, the areas of solid part and working gas are given by
( )
Asolid = A 1 − ψ reg and A fluid = Aψ reg (4.4.15)
Substitution of Eq. (4.4.15) into (4.4.14), it yields
[ ]
Q& cond = − A (1 − ψ reg ) ⋅K s + ψ reg ⋅ K
dTm
dx
. (4.4.16)

By comparison of Eq. (4.4.13) and (4.4.16), the effective thermal conductivity is


( )
K eff = 1 − ψ reg ⋅ K s + ψ reg ⋅ K . (4.4.17)
In this way an estimation of the thermal losses can be made, neglecting any
convective effects and contact to contact thermal resistivities.

4.4.3 Selection criteria


There are three main criteria to choose a regenerator material:
First criterion
The hydraulic radius may not be too large, otherwise the oscillating gas parcels in
the regenerator will not be locally isothermal, nor too small, which could result in
losses due to large fluidic resistance. For the best performance, there should be an
optimum ratio between the hydraulic radius and the working gas’ thermal
penetration depth. Here, the first guess is 1/3 of a reference thermal penetration
depth, i.e. rh ≈ 1 / 3 ⋅ δ gas (the reference thermal penetration depth δ gas is evaluated
with the operation frequency of the system, operation mean pressure and the
working gas property parameters at the average value of the hot and cold-end
temperatures of the regenerator). The specific value of the ratio will be found later
by comparison of the performance of different regenerators. Considering the
correlation between the hydraulic radius and porosity, an optimum value for
porosity is investigated as well.
First, this criterion for hydraulic radius and porosity can be checked by the theory
described in section 4.3. Considering that the experimental set-up for
measurements is a coaxial traveling-wave engine (the details are given in the next
section 4.4.4), the formula for engine mode is used in the following computation.
All the stainless steel screen regenerators, which are measured in this study, are
computed and plotted in the figures. In all computations, the configuration of the
engine, the hot and cold-end temperatures, and the given input acoustic pressure at
Traveling-wave systems 121

some position are the same for all computations of different stainless steel screen
regenerator samples. In other words, the working conditions remain the same for
all computations. Since the experimental set-up is a coaxial traveling-wave engine,
meaning that the geometry is not a clearly defined looped configuration which is
used in section 4.3, the geometrical input parameters are approximated and adapted
to a looped configuration. The thermal penetration depth used as the reference
δ gas for the hydraulic radius in Figs. 4.4.7, 4.4.8 and 4.4.9 is 166 µm. This value is
for a temperature of 490 K of argon, mean pressure of 10 bar and frequency of 59
Hz being the average setting of the traveling-wave system. The efficiency of the
engine as a function of porosity and of the ratio of hydraulic radius and the
reference penetration depth is plotted in Fig. 4.4.7. The graph shows clearly that
there is an optimum in hydraulic radius, and that there is a trend that larger porosity
leads to larger efficiencies.

7.0
6.5
6.0
5.5
5.0
Efficiency [%]

4.5
4.0
3.5
3.0
2.5
2.0
1.5
1.0 0.9
h
0.8
0.7 pt
tio s/

0.5
de
tra diu

0.6
0.0
n

0.0 0.1 0.5


ne r a

0.2 0.4
pe lic

0.3 0.3
al au

0.4
0.5 0.2
rm ydr

Poro 0.6
sity 0.7 0.1
H

0.8 0.0
0.9
e
Th

Figure 4.4.7: Theoretical predictions for stainless steel screen regenerators:


efficiency as a function of porosity and the ratio of hydraulic radius and the
reference thermal penetration depth (i.e.166µm).
122 Chapter 4

4 Porosity 40%
Porosity 60%
Porosity 80%
3
Efficiency [%]

0
0.0 0.2 0.4 0.6 0.8 1.0
Hydraulic radius/Thermal penetration depth

Figure 4.4.8: Theoretical predictions for stainless steel screen regenerators:


efficiency as function of the ratio of hydraulic radius and the reference thermal
penetration depth at fixed porosity.

rh/δgas=0.1
5 rh/δgas=0.2
rh/δgas=0.3
4 rh/δgas=0.5
Efficiency [%]

rh/δgas=0.7

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Porosity

Figure 4.4.9: Theoretical predictions for stainless steel screen regenerators:


efficiency as function of porosity at fixed ratio of hydraulic radius and the
reference thermal penetration depth.

The efficiency at a fixed porosity as function of ratio of hydraulic radius and the
reference penetration depth is given in Fig. 4.4.8. The computation shows that
Traveling-wave systems 123

indeed there exists an optimum value of the ratio for a fixed porosity. It is also seen
that the optimum value for the ratio moves to smaller values when the porosity
becomes larger.
The efficiency at a fixed ratio of hydraulic radius and the reference penetration
depth as function of porosity is given in Fig. 4.4.9. Generally, the performance
increases with porosity. Due to the absence of influence of heat capacity of the
solid part of a regenerator in the model described in section 4.3, the performance
deterioration that must occur at porosities above 90% is not seen in Fig. 4.4.9.
Therefore, the prediction based on the model is not reliable when the porosity is
above 90%.

Second criterion
During the acoustic cycle, heat will penetrate into the regenerator material over a
distance characterized by the solid’s thermal penetration depth. Therefore, it has no
advantage to employ relatively thick wires. The efficiency will decrease if useful
space is wasted by using relatively thick solid material. Therefore, the diameter of
wire δ should be less than twice the solid material’s thermal penetration depth,
δ < 2δ s . (4.4.18)
This criterion will determine the porosity of the regenerator.
Tijani et al. showed that 2yo ≥ 8δs for a plate of stack and maybe also for a wire
[88].

Third criterion
The ratio between the heat capacity of the regenerator material and the heat
capacity of the working gas should be large. A large heat capacity means that a
large amount of energy can be stored per unit volume, which ensures no
temperature oscillation in the solid material to guarantee a stable local temperature.
In the present study, a required ratio of 10 is assumed:
ρ s csVs
> 10 . (4.4.19)
ρ mc pVgas
Vs and Vgas are the volumes of solid material and the working gas.
These criteria are plotted for the honeycomb materials in Figs. 4.4.10 and 4.4.11,
and for the screen materials in Figs. 4.4.12 and 4.4.13. By the three criteria, the
good choices for a regenerator material should be in the range which is below the
dash-dot line (selecton criterion 3), left to the dash line (selecton criterion 2), and
close to the solid line (selecton criterion 1).
124 Chapter 4

selection criterion 1

Open channel diameter [mm]


selection criterion 2
10
1 selection criterion 3
test samples

0
10

-1
10

0 0.05 0.1 0.15 0.2


Wall thickness [mm]
Figure 4.4.10: Selection criteria for ceramic honeycomb regenerators (logarithmic
plot).
Open channel diameter (d) [mm]

selection criterion 1
selection criterion 2
0.5 selection criterion 3
measured honeycombs

0.4

0.3

0.2

0.1

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
Wall thickness (delta) [mm]
Figure 4.4.11: Selection criteria for ceramic honeycomb regenerators (linear
scale).
Traveling-wave systems 125

selection criterion 1
selection criterion 2

Open wire distance [mm]


10
1 selection criterion 3
test samples

0
10

-1
10

0 0.1 0.2 0.3 0.4 0.5


Wire thickness [mm]
Figure 4.4.12: Selection criteria for stainless steel screen regenerators
(logarithmic plot).

0.5
selection criterion 1
Open wire distance (d) [mm]

selection criterion 2
0.4 selection criterion 3
measured screens

0.3

0.2

0.1

0
0 0.1 0.2 0.3 0.4 0.5
Wire thickness (delta) [mm]
Figure 4.4.13: Selection criteria for stainless steel screen regenerators (linear
scale).

4.4.4 Experimental set-up: a coaxial traveling-wave engine


The experimental set-up is shown in Fig.4.4.14. It consists of a coaxial half-wave-
length thermoacoustic engine, a resonator tube, a variable acoustic load, measuring
devices, and data acquisition equipment.
126 Chapter 4

Figure 4.4.14: Schematic view of the coaxial traveling-wave engine test set-up.The
regenerator part is shown in the left-below corner, marked out by an ellipse.

The components are explained below.


The Resonator tube
The resonator tube is made of a symmetric tube of stainless steel which is able to
withstand a pressure up to 20 bar. The dimensions in mm and geometry of the
resonator tube are given in Fig.4.4.15. It is mounted on dampers to minimize the
transfer of vibration. One end is closed by an end flange which is fixed with bolts
and nuts. The thermoacoustic engine is mounted inside of this end. The other end is
empty and connected to a variable acoustic load by a flange.

440 920 440

95.5 54.8 95.5

Figure 4.4.15: Dimensions of the resonator tube (in mm).

Thermoacoustic engine
Fig. 4.4.16 (a) and (b) show the so-called insert of the thermoacoustic engine. From
bottom to top, the engine consists of several elements: membrane, cold-end heat
exchanger (cooler), regenerator, hot-end heat exchanger (heater) and two flow
Traveling-wave systems 127

straighteners. Two cooling liquid flow tubes connect the heat exchanger to a
thermal Julabo cooler, as seen in Fig. 4.4.16 (a). A thermal insulation ring of
Vespel surrounds the cooler, regenerator and heater. It is placed between the holder
tube and the part consisting of cooler, regenerator and heater. The insulation ring
has a thickness of 0.52 mm and is able to withstand a temperature up to 400 °C.
The holder tubes are supported by three identical threaded rods. The rim of the
holder of the thermoacoustic engine is mounted at a distance of 130 mm from the
end flange of the resonator tube.

The elements are described in detail below.


Membrane: This is a latex diaphragm closing the lower part of the engine to
prevent Gedeon streaming. This streaming refers to a non-zero time average mass
flow in the toroidal space (around the cylinder formed by the regenerator and the
heat exchangers in this experiment set-up) shuttling heat from hot end to the cold
end. The membrane allows the gas to oscillate, but blocks the net mass flow at this
position.

(a) Photo of the engine.


The upper part points to the
(b) Sketch of the engine and all the components
load side of the resonator
with their positions in the tube.
tube.
Figure 4.4.16: The coaxial thermoacoustic engine.
128 Chapter 4

(a) Photo of the cold-end heat exchanger (b) Photo of the hot-end heat exchanger
(cooler) (heater)

Figure 4.4.17: The heater and cooler of the thermoacoustic engine.

Cold-end heat exchanger (cooler): The temperature of the cold end of the
regenerator is fixed by a heat exchanger with cooling water. The photo is given in
Fig. 4.4.17 (a). A Julabo HL-4F32 cooler is used to maintain the temperature of the
water at a constant temperature of 22 °C.
Regenerator: The regenerator is a wire-screen filled cylinder with a height of 11.2
mm and a diameter of 55.0 mm. Wire screens are stacked up to fill the available
space. All the tested samples used in this experiment fit in this space.
Hot-end heat exchanger (heater): The engine is driven by heat generated with an
electric heater. A spiral shaped wire heater is mounted on a perforated plate. The
heater is connected to an electric power supply from Delta Elektronica type
SM3004-D. The power is varied to heat the hot end of the regenerator to a
maximum temperature of 450°C. The heater is shown in Fig. 4.4.17 (b).
First flow straightener: A first flow straightener is positioned directly above the
heater. It is made from porous metallic material and used to make the acoustic flow
uniform.
Second flow straightener: A second flow straightener is mounted 10 cm further in
the heater tube. It functions similarly as the previous one, but with larger pores.

The variable acoustic load


A variable acoustic load is connected to the end of the resonator tube. It consists of
a throttling valve and a buffer vessel with a volume of 1 dm³, as schematically
given in Fig. 4.4.18. The dissipated acoustic power can be measured by comparing
the phase and amplitude of the acoustic wave at the end of the resonator with those
in the buffer vessel.
Traveling-wave systems 129

AC load

Figure 4.4.18: The variable acoustic load with pressure sensors.

4.4.5 Energy balance in the experimental set-up

Figure 4.4.19: The energy balance for the thermoacoustic engine. The energy flow
out of the box 1 equals to the flow into box 2.

The energy balance of the thermoacoustic engine is shown in Fig. 4.4.19. In this
experimental setup, the heat to drive the engine is supplied by an electrical heater.
This gross energy is partially lost to the environment through the walls of the
resonator tube, and only a fraction is used by the engine to convert heat to acoustic
energy. The whole system is approximately symmetric. Thus, if the system is
divided into two parts at the middle of the resonator tube, the amount of acoustic
power dissipated in one part approximately equals that in the other part, as seen in
Fig. 4.4.19. The dissipated acoustic power, Pdissipated , is due to thermoviscous losses
by motion of gas near the walls of the tube and radiation out of the tube. The

acoustic energy Pacoustic that leaves the left confined imaginary box 1 is:
− 1
Pacoustic = Pheater − Pcooler − Plosses − Pdissipated . (4.4.20)
2
+
The acoustic energy Pacoustic into the right part confined imaginarily by box 2 is
+ 1
Pacoustic = Pload + Pdissipated . (4.4.21)
2
130 Chapter 4

Pacoustic is the acoustic power measured at the center of the resonator tube.
According to energy conservation, the energy that flows out of box 1 equals that
entering box 2:
+ −
Pacoustic = Pacoustic = Pacoustic . (4.4.22)
The acoustic energy dissipated in the variable load Pload is in principle useful energy
that can be applied as power for instance to convert into mechanical energy.
All energy losses in the engine section are counted into Plosses . They are heat
convection from the hot end of the regenerator to the cold end without any acoustic
power production, thermal radiation losses, as well as Gedeon streaming (“DC”
flow).

Performance of the engine


The performance of an engine can be evaluated by the raw efficiency η or the
fraction of the Carnot efficiency ηrelative ( ηrelative = η / ηCarnot ). The raw efficiency
η is the ratio of useful work, produced by the engine, to the total heat added to the
system. In this case, the energy Pload is taken as useful work, produced by the
system, in the consideration that the dissipated acoustic power along the
walls, Pdissipated , is unavailable for practical use. Thus, the raw efficiency is
calculated by
η = Pload / Pheater × 100% . (4.4.23)

4.4.6 Measurement equipment and data handling


The main concerns of this measurement are the energy flows. To determine those,
pressure sensors and temperature sensors are mounted. All the measuring devices
and the connections to the computer are schematically shown in Fig. 4.4.20. There
are six temperature sensors in the system. Two Pt1000 thermometers are used to
measure the temperatures of the cooling water flowing in and out of the cooler,
respectively. Three thermocouples are placed in the regenerator, one at the hot end,
the second at the middle, and the third at the cold end, as the enlarged section of
Fig. 4.4.20 shows. One thermocouple is placed between two flow straightners
outside the regenerator in the thermal buffer tube zone. Six pressure sensors are
mounted in the system. Three microphones are mounted in the wall around the
middle part of the resonator tube to acquire the information for calculating Pacoustic .
Two microphones are mounted around the variable acoustic load to calculate the
acoustic power dissipated inside. One pressure sensor is mounted on the surface of
the end flange connecting to the variable acoustic load, which is for the static mean
pressure measurement.
Traveling-wave systems 131

Figure 4.4.20: Overview of the measuring devices and data acquisition.

During the measurement, all data are collected with a National Instruments PXI-
system and handled by a LabView program. The parameters are monitered via a
LabView data acquisition system. The system composes two files for every
measuring point. These data files are analyzed with a Matlab data processor. In the
following, the measurement of every unit of the thermoacoustic system is discussed,
as well as the calculation of energy flows.

Cooling and heater power


The supplied heat from the electrical heater is calculated by
Pheater = U ⋅ I , (4.4.24)
where U and I are the voltage and current of the power supply, respectively.
The waste heat dumped at the cold end of the regenerator is removed by the
cooling water through the cooler. The amount of heat Pcooler is calculated by
Pcooler = V& ⋅ ρ water ⋅ c pwater ⋅ (Tout − Tin ) , (4.4.25)
&
where V is the flow rate of the cooling water. Tin and Tout are the temperatures of
the cooling water flowing in and out from the cooler, and are measured by two
thermocouples. ρ water and c pwater are density and specific heat of water,
respectively.
Acoustic energy
The acoustic intensity is determined from the signals of two adjacent pressure
sensors, by using the average of the two signals to obtain the pressure and the
difference between the two signals to obtain velocity. It is called the “Two-
microphone method”, which is described in reference [62]. Three sensors are
132 Chapter 4

applied for generating three acoustic flow measurements. The signals of the three
pressure sensors mounted halfway of the resonator tube are used to obtain the
acoustic power following:
⋅ pB ⋅ pC ⋅ sin (φB − φC ) ,
A
Pacoustic = (4.4.26)
2ωρ gas ∆x
where A is the cross-sectional area of the resonator tube, and B,C are two points
with a distance of ∆x along the tube. pB and pC are the pressure amplitudes at
points B and C respectively. φB and φC are the phase angles of the pressure at
these points.
Power to the acoustic load
The power delivered to the variable acoustic load is determined by using a similar
method as in reference [62]. Information from two microphones, one upstream of
the throttling valve and the other in the buffer, is used. It is schematically
illustrated in Fig.4.4.18. The energy dissipated by the throttling valve is given by
1 ωV0
Pload = ⋅ ⋅ pE ⋅ pT ⋅ sin (φE − φT ) , (4.4.27)
2 γ P0
where γ is the ratio between the isobaric and isochoric specific heats and V0 is the
volume of the buffer vessel. P0 is the mean pressure in the tube, measured by a
static pressure sensor at the end of the tube in form of relative pressure compared
to the atmospheric pressure. A dynamic pressure sensor upstream of the throttling
valve measures the pressure amplitude pE and phase angle φE . Similarly, another
dynamic pressure sensor in the buffer vessel measures the pressure amplitude pT
and phase angle φT .
Dissipated energy
Pdissipated is the amount of acoustic energy that is dissipated along the walls of the
resonator tube. If the acoustic energy Pacoustic and the energy to the load Pload are
determined as described above, the dissipated energy can be calculated by Eq.
(4.4.21)
Pdissipated = 2 ⋅ (Pacoustic − Pload ) . (4.4.28)
Energy losses in TA-engine
The lost energy in the engine is mostly due to convection and radiation to the
surroundings, as shown in Fig.4.4.19. With all the energy terms known by above
calculations, the energy Plosses can be determined by using Eq. (4.4.20)
1
Plosses = Pheater − Pcooler − Pacoustic − Pdissipated . (4.4.29)
2
Traveling-wave systems 133

4.4.7 Measuring procedure


Preparation of the set-up
For every measurement run, only one regenerator sample is tested. After the
regenerator is placed in the holder and the system is closed, the system is pumped
off by a rotation pump and an oil-diffusion pump. Then, the system is filled with
the desired gas, helium or argon at an absolute pressure of 10 bar. While filling, the
pressure can be read on the Labview window.
Starting up
The system produces stable acoustic resonance only if the temperature gradient is
large enough, which means that the heat, fed by the heater, must exceed some
threshold. In this set-up, the heater power can be varied by changing the voltage or
current of the electric power supply. When the engine starts to generate sound and
a stable resonance is produced, it can be seen on the oscilloscope visualization of
the pressure signals. Then, the XY-visualization of the two outer microphones at
the center of the resonator tube shows a stable ellipse and all the five dynamic
pressure sensors show harmonic oscillation. At the point of threshold it is often
observed that the sound intensity increases and, while the thermoacoustic heat
pumping starts, the thermal balance in the regenerator is disturbed. This leads to an
unstable mode with a periodicity of several minutes. So, when onset is reached,
sound generation occurs, heat pumping starts, and because of the thermal balance
in the regenerator being disturbed, the sound production decays to zero. Over a
window of several watts, this unstable behavior occurs, but finally at a sufficiently
large heat input a stable sound intensity is maintained. In the starting-up stage, no
net acoustic energy is produced and the wave is not stable. Temperature oscillation
occurs during this stage. De Waele [70] gave a detailed theoretical analysis on this
onset temperature and oscillation stability of a traveling-wave engine with a torus-
shaped section. Similar results were obtained in LeMans at the laboratoire
acoustique, see Penelet [87].
Measuring
To start the measurements of the performance of the engine, normally, a power of
10 W is fed to the heater. The temperature evolution can be followed on the
Labview program window. After all the temperatures have reached stable values,
data from all the sensors and thermocouples, as well as information for the heater
and cooling water flow are stored in two files. For about thirty seconds, all the data
are stored. Then, heater power is stepwise increased to start the next measurement.
If the engine starts to produce sound, the throttling valve for acoustic load has to be
opened to maintain a fixed drive ratio of 2.7%. The drive ratio is the ratio between
the amplitude of the sound wave at the end of the resonator tube and the mean
pressure inside the system. When more and more heat is added to the heater, the
134 Chapter 4

throtting valve must be opened more and more correspondingly to maintain a


constant drive ratio. The steps are repeated for every energy input to the heater till
the maximum temperature of the hot end of the regenerator reaches around 450 °C,
which is the maximum temperature allowed. All the files are processed using a
Matlab program to compute all the parameters of interest as a function of
increasing Pheater , or as a function of the corresponding temperature difference
across the regenerator. The data of one measurement of 30 seconds are processed
by the Matlab program to one average value as the value for analysis.
List of standard measurement settings
The measurements are conducted under standardized measurement settings needed
for calculations, as listed below:
Working gas: Argon or helium;
Mean pressure: 10 bar;
Temperature of cooling water flowing in the cold end heat exchanger: 22 °C;
Drive ratio: 2.7%;
Maximum temperature at hot end of the regenerator: 720 K;
Mean regenerator gas temperature used for calculations: 490 K, which is an
estimate of the mean temperature during a measurement. This temperature, 490 K,
is also used in the tables below as average temperature to determine the properties
of the gas, and related dimension(less) numbers.

Regenerator samples
The ceramic honeycomb regenerators were supplied by Corning and the metal
honeycomb samples were manufactured at the University of Liverpool. The
dimensions of all honeycomb regenerators were measured using a microscope. The
wire screens are from Metaalgaasweverij Dinxperlo. The mesh number M and
diameter of wires δ are listed by the manufacturer and used for calculations.
Values of hydraulic radius rh , hydraulic diameter Dh , and porosity ψ reg are
calculated by corresponding equations in section 4.4.2. The basic average gas’
thermal penetration depth δ κ average , which works as reference in later
computations is calculated by using the properties of the working gas at 10 bar and
490 K. For argon, δ κ average is 166 µm. Some important parameters of all the
samples, which are tested in this study, are listed in the table 4.4.I and 4.4.II.
Traveling-wave systems 135

δ
CPSI M δ Tg λeff
regenerator type gas d [#cells
rh Dh ψ reg
[mm] [mm] [#cells/inch ] 2
/inch] [µm] [µm] rh [W/mK]
ceramic honeycomb Ar 0.08 (±0.01) 0.55 (±0.01) 1600 40 137.5 550 0.76 1.2 0.5
ceramic honeycomb Ar 0.06 (±0.02) 0.43 (±0.02) 2700 52 107.5 430 0.77 1.6 0.48
ceramic honeycomb Ar 0.07 (±0.01) 0.34 (±0.01) 3800 62 85 340 0.69 2.0 0.64
ceramic honeycomb Ar 0.06 (±0.01) 0.28 (±0.01) 7800 75 70 280 0.68 2.4 0.66
ceramic honeycomb Ar 0.03 (±0.01) 0.09 (±0.01) 43300 208 22.5 90 0.56 7.4 0.89
ceramic honeycomb Ar 0.03 (±0.02) 0.04 (±0.02) 106000 326 10 40 0.33 16.7 1.36
ceramic honeycomb Ar 0.02 (±0.01) 0.08 (±0.01) 46300 215 20 80 0.64 8.3 0.74
ceramic honeycomb Ar 0.02 (±0.01) 0.13 (±0.01) 30300 174 32.5 130 0.75 5.1 0.52
metal honeycomb Ar 0.12 (±0.01) 0.63 (±0.01) 1100 33 157.5 630 0.71 1.1 18
metal honeycomb Ar 0.12 (±0.01) 0.38 (±0.01) 2600 51 95 380 0.58 1.8 25
metal honeycomb Ar 0.12 (±0.01) 0.28 (±0.01) 4000 63 70 280 0.49 2.4 31
metal honeycomb He 0.12 (±0.01) 0.63 (±0.01) 1100 33 157.5 630 0.71 1.7 18
metal honeycomb He 0.12 (±0.01) 0.28 (±0.01) 4000 63 70 280 0.49 3.9 31

Table 4.4.I Ceramic and metal honeycomb regenerator properties.

CPSI M δ Tg λeff
regenerator type gas δ d
[#cells/i [#cells/ γ pf rh Dh ψ reg
[mm] [mm] 2 [µm] [µm] rh [W/mK]
nch ] inch]
stainless steel wire screeens Ar 0.035-0.036 0.066 62500 250 1.8 20.3 81.2 0.70 8.17 8.23
stainless steel wire screeens Ar 0.035-0.036 0.059 72900 270 2.0 20.8 83.1 0.70 7.99 8.10
stainless steel wire screeens Ar 0.035-0.036 0.075 52900 230 1.9 25.0 99.8 0.74 6.65 7.10
stainless steel wire screeens Ar 0.035-0.036 0.049 90000 300 1.6 12.5 50.1 0.59 13.26 11.22
stainless steel wire screeens Ar 0.05 0.104 27225 165 2.1 38.6 154.3 0.76 4.30 6.63
stainless steel wire screeens Ar 0.03 0.224 10000 100 1.6 58.2 232.8 0.89 2.85 3.11
stainless steel wire screeens Ar 0.03 0.288 6400 80 1.6 72.8 291.3 0.91 2.28 2.55
stainless steel wire screeens Ar 0.03 0.478 2500 50 1.3 93.6 374.3 0.93 1.77 2.03
stainless steel wire screeens Ar 0.1 0.218 6400 80 1.8 65.5 261.8 0.72 2.54 7.48
stainless steel wire screeens Ar 0.1 0.147 10609 103 1.7 39.8 159.0 0.61 4.18 10.44
stainless steel wire screeens Ar 0.06 0.115 21025 145 2.1 44.4 177.5 0.75 3.74 6.84
stainless steel wire screeens Ar 0.05 0.077 40000 200 1.9 25.1 100.4 0.67 6.61 8.99
stainless steel wire screeens Ar 0.039-0.040 0.102 32400 180 1.9 32.9 132.0 0.77 5.06 6.26
stainless steel wire screeens Ar 0.039-0.040 0.087 40000 200 1.9 28.4 114.0 0.74 5.87 6.99
stainless steel wire screeens Ar 0.039-0.040 0.062 62500 250 1.9 20.8 83.3 0.68 8.00 8.70
stainless steel wire screeens Ar 0.035 0.110 32400 180 2.0 37.2 149.0 0.81 4.48 5.17

Table 4.4.II Stainless steel wire screens regenerator properties.

4.4.8 Results and discussion


The results of measurements on the performance of different regenerator samples
are presented in this subsection. All data are obtained at operation under the
standard measurement settings described in the last subsection, except for some
samples. One exception is the sample of ceramic honeycomb regenerator with a
CPSI of 43300. It was measured twice: one time without the outer flow straightner
for comparison with standard condition, which is indicated by (1) in Fig. 4.4.26,
and the other one with both flow straightners as in standard condition, which is
indicated by (2). Another exception is about the stainless steel wire screen
regenerator with mesh 180 and wire diameter 0.035-0.036 mm. It was also
measured twice: one time, indicated as (1), was measured at a lower drive ratio of
2.3% than the standard one, and the second time, indicated as (2), was measured at
standard drive ratio of 2.7%.

Energy balance
According to the energy equations (4.4.20) (4.4.21) and (4.4.22), the ratios of four
energy consumptions to total heater power have to sum up to 1:
136 Chapter 4

Pcooler Plosses Pacoustic Pdissipated / 2


+ + + = 1. (4.4.30)
Pheater Pheater Pheater Pheater
In Fig.4.4.21-4.4.23, the relative fraction of the power contribution to the following
four quantities is plotted: heat power transported to the cold end of the regenerator
Pcooler (blue), heat losses to the environment Plosses (green), acoustic power in the
valve Pacoustic (yellow), and acoustic dissipation Pdissipated/2 (pink). It is clear that the
blue zone in all the graphs is dominant. Except for the metal honeycombs, it
becomes in general smaller at increasing heat power. But at the threshold when
acoustic power starts to be generated there is an increase in cooler losses, which is
due to increase of thermal transport in the regenerator due to the acoustics. At
increasing heat power the green zone increases, which is due to increased
convective losses and thermal radiation. Both effects are nonlinear and although
the fractional losses are plotted, these losses contribute more and more at the
expense of the cooler losses. The yellow and pink zones are present only after the
onset of the acoustic wave. The fraction of the acoustic power and losses is small
with respect to the other two losses. This is mainly due to two aspects, the coaxial
geometry is not insulated and there is a large thermal gradient over a thin
regenerator (about 10 mm thick) leading to significant thermal radiation and
conduction losses. In case of the metal honeycomb, see Fig. 4.4.21, the production
of acoustic power remains very small due to the relatively large size of the
channels compared to the penetration depth as well as to the huge axial heat
conduction. Note that the x axis scale extends here to 270 W still with negligible
acoustic power production. The poor performance of metallic honeycombs is also
clearly shown in Figs. 4.4.24 and 4.4.25. The production of Pacoustic as a function of
∆T (∆T=TH-TC, where TH and TC are the temperatures of hot and cold-end of the
regenerator, respectively) for metallic honeycombs is given in Fig. 4.4.26.
From the measurement of the temperature difference over the regenerator, we
conclude that in principle at a ∆T of 80 K, the engine already produces acoustic
power. In most of the measurements with metal honeycombs, this ∆T is attained.
However, the cell size of these honeycombs is too large (too small CPSI number).
Only for CPSI number of 4000 there is acoustic power generated. Of course,
acoustic power can be generated with metal honeycombs of CPSI=1100 or 2600,
however only at powers and temperatures far above the available power that we
have. The main reason for not generating acoustic power is the disagreement
between cell size and thermal penetration depth.
Traveling-wave systems 137

100

80

Power distribution [%]


60

40
Pcooler/Pheater [%]
Plosses/Pheater [%]
20 Pacoustic/Pheater [%]
(Pdissipated/2)/Pheater [%]

0
0 25 50 75 100 125 150 175 200 225 250 275
Pheater [W]

Figure 4.4.21: Relative power distribution for the metal honeycomb regenerator of
CPSI 4000.

100

80
Power distribution [%]

60

Pcooler/Pheater [%]
40 Plosses/Pheater [%]
Pacoustic/Pheater [%]
(Pdissipated/2)/Pheater [%]
20

0
0 10 20 30 40 50 60 70 80 90 100 110
Pheater [W]

Figure 4.4.22: Relative power distribution for the ceramic honeycomb regenerator
of CPSI 46300.
138 Chapter 4

100

80
Power distribution [%]

60

40 Pcooler/Pheater [%]
Plosses/Pheater [%]
Pacoustic/Pheater [%]
20 (Pdissipated/2)/Pheater [%]

0
0 10 20 30 40 50 60 70 80 90
Pheater [W]

Figure 4.4.23: Relative power distribution for the stainless steel screen
regenerator with mesh 180 and wire diameter 0.035.

CPSI 1100, Ar
2.5
CPSI 1100, He
CPSI 2600, Ar
2.0 CPSI 4000, Ar
CPSI 4000, He(1)
CPSI 4000, He(2)
1.5
Pacoustic (W)

1.0

0.5

0.0

-0.5
0 50 100 150 200 250 300 350 400
Pheater (W)

Figure 4.4.24: Pacoustic as a function of Pheater for metal honeycomb regenerators.


Traveling-wave systems 139

0.20 CPSI 1100, Ar


CPSI 1100, He
CPSI 2600, Ar
CPSI 4000, Ar
0.15
CPSI 4000, He(1)
CPSI 4000, He(2)

0.10
Efficiency [%]

0.05

0.00

-0.05
0 50 100 150 200 250 300 350 400
Pheater (W)

Figure 4.4.25: Efficiency as a function of Pheater for metal honeycomb regenerators.

2.5

CPSI 1100, Ar
CPSI 1100, He
2.0 CPSI 2600, Ar
CPSI 4000, Ar
CPSI 4000, He(1)
1.5 CPSI 4000, He(2)
Pacoustic (W)

1.0

0.5

0.0
0 20 40 60 80 100 120 140 160 180
∆T (K)

Figure 4.4.26: Pacoustic as a function of ∆T for metal honeycomb regenerators.

Power dissipated in the variable acoustic load


The energy, dissipated in the acoustic load, is determined via Eq. (4.4.27). The
energy of Pload as a function of the heating power Pheater for ceramic honeycomb
samples and stainless steel screen regenerators is plotted in Figs. 4.4.27 and 4.4.28.
140 Chapter 4

The production of Pacoustic as a function of ∆T for ceramic honeycombs and


stainless steel screen are given in Fig. 4.4.29 and 4.4.30.
For an ideal engine that operates under ideal circumstances with no losses, it can be
loaded as soon as any heater energy is applied, so Pload increasing fundamentally
linear with Pheater . However, in a real engine, there exists a threshold value for
Pheater below which there is no acoustic energy production. From Eq. (4.4.21),
power can only be added to the acoustic load if Pacoustic > Pdissipated / 2 . Below that,
the efficiency is zero and no useful energy is provided by the engine. Under those
heat loads, all power produced by the engine is used up by losses and dissipation
along the walls. For most of the ceramic samples, this threshold point is at
Pheater ≈ 40 − 50W . For stainless steel screen samples, the point is reached at 20-
30 W. The three ceramic samples, with CPSI lower than 7800, do not give any
significant Pload . Here the thermal penetration depth of the wave is much smaller
than the characteristic cell size of the material. For the ceramic honeycomb
regenerators, the one with CPSI 30300 gives the maximum of Pload = 3.5W at
Pheater = 130W . It is clear from Fig. 4.4.27 that the threshold and performance of
the honeycombs is related to the CPSI number or effective cell width.
In general, the stainless steel screen samples generate more power and perform
better than the ceramic honeycombs. For the stainless steel screen samples, the one
with mesh 80 and wire diameter of 0.03 mm gives the maximum of
Pload = 6.7 W at Pheater = 140W . But the best performance in terms of efficiency is
with mesh 100/0.03mm.
In these two figures, the curves are almost linear after the threshold points. The
sample of mesh 80 and wire diameter of 0.03 mm in Fig.4.4.28, for instance, shows
this typical trend. As stated above, in ideal situation, Pload increases linearly
with Pheater , Pload , ideal = α load ⋅ Pheater . In real situation, the acoustic power is
generated after Pheater exceeds the threshold value Pthreshold . As seen in Fig. 4.4.29,
at this threshold power, the ∆T for these materials is approximately 85 (±10) °C.
Physically speaking this could be defined as the lowest threshold ∆T at optimum
CPSI number for this engine. As seen in Figs. 4.4.27 and 4.4.28, Pload increases
linearly with Pheater after reaching the threshold value:
Pload = α load ⋅ (Pheater − Pthreshold ) .
In Fig.4.4.30, similar to Fig.4.4.28, Pacoustic increases almost linearly with ∆T after
reaching the threshold ∆T for a specific stainless steel screen material
(distinguished by mesh number and wire diameter).
Traveling-wave systems 141

CPSI 1600
CPSI 2700
CPSI 3800
CPSI 7800
4.0 CPSI 43300(1)
CPSI 106000
3.5
CPSI 46300
CPSI 30300
3.0
CPSI 43300(2)
2.5

2.0
Pload (W)

1.5

1.0

0.5

0.0

-0.5
0 20 40 60 80 100 120 140 160 180 200 220 240
Pheater (W)

Figure 4.4.27: Pload as a function of Pheater for ceramic honeycomb regenerators.

6
mesh180,wire0.035-0.036(1)
mesh180,wire0.039-0.040
5
mesh200,wire0.039-0.040
mesh250,wire0.039-0.040
4
mesh180,wire0.035-0.036(2)
Pload (W)

mesh250,wire0.035-0.036
3 mesh230,wire0.035-0.036
mesh165,wire0.05
2 mesh270,wire0.035-0.036
mesh300,wire0.035-0.036
1 mesh100,wire0.03
mesh80,wire0.03
0 mesh50,wire0.03

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150


Pheater (W)

Figure 4.4.28: Pload as a function of Pheater for stainless steel screen regenerators.
142 Chapter 4

CPSI 1500
6
CPSI 3000
CPSI 3900
5 CPSI 7300
CPSI 43300(1)
CPSI 106000
4 CPSI 46300
CPSI 30300
Pacoustic (W)

CPSI 43300(2)
3

0
0 50 100 150 200 250 300 350 400
∆T (K)

Figure 4.4.29: Pacoustic as a function of ∆T for ceramic honeycomb regenerators.

mesh180,wire0.035-0.036(1)
10 mesh180,wire0.039-0.040
mesh200,wire0.039-0.040
9
mesh250,wire0.039-0.040
8 mesh180,wire0.035-0.036(2)
mesh250,wire0.035-0.036
7 mesh230,wire0.035-0.036
mesh165,wire0.05
6
mesh270,wire0.035-0.036
Pacoustic (W)

5 mesh300,wire0.035-0.036
mesh100,wire0.03
4 mesh80,wire0.03
mesh50,wire0.03
3

0
0 50 100 150 200 250 300 350 400 450
∆T (K)

Figure 4.4.30: Pacoustic as a function of ∆T for stainless steel screen regenerators.

Performance of the engine


Efficiency
By using Eq. (4.4.23), the raw efficiencies of ceramic honeycomb and stainless
steel screen regenerators are determined and plotted against Pheater , as shown in Figs.
Traveling-wave systems 143

4.4.31 and 4.4.32. As seen in the discussion of Pload , there is a threshold value
below which there is no energy production. After that, the efficiency increases as
the heater power increases. For some samples, the efficiency increases to a
maximum value and then slightly decreases. The decrease after reaching a
maximum could be possibly attributed to the increasing nonlinear convection
losses.
The effect of a second flow straightner can be found in the comparison between
two measurements of the ceramic honeycomb sample with CPSI 43300, as seen in
Fig. 4.4.31. Apparently, the performance using two flow straightners, marked with
(2) in the figures, is better than the one without second flow straightner, marked
with (1). It can be concluded that the second flow straightner prevents convection
losses. Similarly, two working conditions were applied to the stainless steel screen
sample with mesh 180 and wire diameter of 0.035 mm. One time, the regenerator
worked at drive ratio of 2.3%, marked by (1), and the second time, at standard
drive ratio of 2.7%, marked by (2). The results show that the higher drive ratio
gives a much better performance.

CPSI 1600
CPSI 2700
3
CPSI 3800
CPSI 7800
CPSI 43300(1)
CPSI 106000
CPSI 46300
2 CPSI 30300
CPSI 43300(2)
Efficiency [%]

0
0 20 40 60 80 100 120 140 160 180 200 220 240
Pheater(W)

Figure 4.4.31: Efficiency as a function of Pheater for ceramic honeycomb


regenerators.
144 Chapter 4

mesh180,wire0.035-0.036(1)
6 mesh180,wire0.039-0.040
mesh200,wire0.039-0.040
5 mesh250,wire0.039-0.040
mesh180,wire0.035-0.036(2)
mesh250,wire0.035-0.036
Efficiency [%]
4
mesh230,wire0.035-0.036
mesh165,wire0.05
3 mesh270,wire0.035-0.036
mesh300,wire0.035-0.036
2
mesh100,wire0.03
mesh80,wire0.03
mesh50,wire0.03
1

0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150
Pheater (W)

Figure 4.4.32: Efficiency as a function of Pheater for stainless steel screen


regenerators.

For ceramic honeycomb regenerators, the highest efficiency reached by the


samples is 2.6%, whereas for stainless steel screen regenerators, the maximum is
5.8% (sample with mesh 100 and wire diameter of 0.03 mm). Generally, the
stainless steel screen regenerators give higher efficiency than the ceramic
honeycombs.
Efficiency relative to the Carnot efficiency
The relative efficiency is evaluated as the ratio of raw efficiency to the Carnot
effeciency ηrelative ( ηrelative = η / ηCarnot ). The Carnot efficiency is the theoretical
maximum that an engine is able to reach and that is bound by the second law of
thermodynamics:
TH − TC
ηCarnot = , (4.4.31)
TH
where TH and TC are temperatures at the hot and cold ends.When the heating power
Pheater increases, the Carnot efficiency increases because of the increasing
temperature difference across the regenerator.
The relative efficiency against Pheater for both kinds of samples are plotted in Figs.
4.4.33 and 4.4.34. The maximum relative efficiency for the ceramic honeycomb
samples is around 7%, (sample with a CPSI of 30300 and 46300). Again, the
stainless steel screen samples give better performance than the ceramic
honeycombs. The maximum relative efficiency for the stainless steel screen
samples is 11.7% (sample with mesh 100 and wire diameter of 0.03 mm).
Traveling-wave systems 145

8
CPSI 1600
CPSI 2700
CPSI 3800
7
CPSI 7800
CPSI 43300(1)

Efficiency/Carnot efficiency [%]


6
CPSI 106000
CPSI 46300
5
CPSI 30300
CPSI 43300(2)
4

0
0 20 40 60 80 100 120 140 160 180 200 220 240

Pheater(W)

Figure 4.4.33: Relative efficiency as a function of Pheater for ceramic honeycomb


regenerators.

mesh180,wire0.035-0.036(1)
13 mesh180,wire0.039-0.040
Efficiency/Carnot efficiency [%]

12 mesh200,wire0.039-0.040
11 mesh250,wire0.039-0.040
mesh180,wire0.035-0.036(2)
10
mesh250,wire0.035-0.036
9 mesh230,wire0.035-0.036
8 mesh165,wire0.05
7 mesh270,wire0.035-0.036
mesh300,wire0.035-0.036
6
mesh100,wire0.03
5 mesh80,wire0.03
4 mesh50,wire0.03
3
2
1
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150
Pheater (W)

Figure 4.4.34: Relative efficiency as a function of Pheater for stainless steel screen
regenerators.

The maximum relative efficiency points out the optimum value for Pheater . For the
ceramic samples, the optimum Pheater is between 70-90 W, and 60-90 W is optimum
for the stainless steel screen samples.
146 Chapter 4

Maximum efficiency as a measure for the performance


The overall machine efficiency is low and less than 10%, and even the relative
efficiency is not over 13%. For the main goal of this work, namely to compare
different regenerator materials, this is not a problem. This study aimed at finding
the optimum geometrical properties for a regenerator material in a thermoacoustic
system with respect to the cell size or mesh number related to the thermal
penetration depth. Based on measurements, stainless steel wire screen regenerators
give the best performance among the three kinds of samples tested in this study.
For a regenerator, many geometrical parameters, such as CPSI number, hydraulic
radius, porosity, mesh number as described in section 4.4.2 are used for
characterizing them, and they are inter-related. Among these parameters, the
porosity and hydraulic radius are chosen for this study, which are direct for
analysis and include inherently other parameters by mutual dependency.
The maximum raw efficiency is plotted based on experimental measurements as
functions of porosity and the dimensionless hydraulic radius (ratio of hydraulic
radius and thermal penetration depth at 490 K of argon, mean pressure of 10 bar
and frequency of 59 Hz ,which is 166 µm), respectively.
From the measurements, there exists an optimum value for dimensionless hydraulic
radius (the ratio of hydraulic radius to thermal penetration depth at average
temperature 490 K), which is around 0.3 for stainless steel screens, as seen in Fig.
4.4.35. The trend for stainless steel screens agrees with the computation in Fig.
4.4.8. Although there are some apparently unexpected low values, for instance at
0.4 (mesh 80 and wire diameter 0.1 mm), but the trend is clear. The optimum range
is obvious. Those unexpected points diverging from the trend can possibly be
attributed to the effect of porosity. This becomes clear when interpreting Fig.
4.4.36 where porosity is plotted as a function of dimensionless hydraulic radius.
The low efficiency results in Fig. 4.4.35 correspond to the low porosity samples in
Fig. 4.4.36 as discussed. So this shows experimentally that both porosity and
hydraulic radius are vital parameters. If both porosity and dimensionless hydraulic
radius are used as inputs in the model in section 4.4.3, the computation of stainless
steel wire screen regenerators also gives similar distribution as in the experiment.
This is shown in Fig. 4.4.37.
For ceramic honeycomb samples, the optimum value is much lower than that of
stainless steel screens, which is in the range 0.14-0.2 as shown in Fig. 4.4.35.
Traveling-wave systems 147

Maximum efficiency [%]


4

stainless steel wire screens


1 ceramic honeycomb

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Hydraulic radius/Thermal penetration depth

Figure 4.4.35: Measured maximum efficiency as a function of dimensionless


hydraulic radius for stainless steel screen regenerators and ceramic honeycombs.

1.0

0.9

0.8

0.7

0.6
Porosity

0.5

0.4

0.3

0.2

0.1

0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Hydraulic radius/Thermal penetration depth

Figure 4.4.36: Porosity as a function of dimensionless hydraulic radius for


stainless steel screen regenerators.
148 Chapter 4

Maximum efficiency [%] 5

1 experiments
computation
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Hydraulic radius/Thermal penetration depth

Figure 4.4.37: Computed maximum efficiency as a function of dimensionless


hydraulic radius for stainless steel screen regenerators.

6
stainless steel wire screens
ceramic honeycomb
5
Maximum efficiency [%]

0
0 0.2 0.4 0.6 0.8 1
Porosity

Figure 4.4.38: Measured maximum efficiency as a function of porosity for stainless


steel screen regenerators and ceramic honeycomb.
Traveling-wave systems 149

6 experiments
computation

Maximum efficiency [%]


5

0
0 0.2 0.4 0.6 0.8 1
Porosity

Figure 4.4.39: Computed maximal efficiency as a function of porosity for stainless


steel screen regenerators.

The measured maximum efficiency against porosity is given in Fig. 4.4.38. The
same trend was already explained in Fig. 4.4.9. The efficiency monotonically
increases with porosity, except for very high porosity when the heat capacity of the
solid part of the regenerator is not big enough to sustain a stable local temperature.
The dip at around porosity 0.7 can be explained by the variation of dimensionless
hydraulic radius. In the experiments, the variation of porosity is accompanied by
the variation of hydraulic radius. Therefore, the diverging points from the general
trend emerge as a result of the combined variation of porosity and hydraulic radius.
If both porosity and dimensionless hydraulic radius are used as inputs in the
computation as discussed in section 4.4.3, the computation also gives similar
distribution as in the experiment, which is shown in Fig. 4.4.39.

4.4.9 Conclusion
The performance of a traveling-wave thermoacoustic system is affected by the
geometrical properties of the regenerator in a profound manner.
From this study, it can be concluded that there is an optimum value for the
dimensionless hydraulic radius, which is around 0.3 for stainless steel wire screen
regenerators and 0.14-0.2 for square cell honeycombs. The efficiency goes up with
increasing porosity, to a maximum value above which the heat capacity of the solid
becomes the limiting factor.
150 Chapter 4

The maximum efficiency that could be established with regenerator in experiments


corresponded well with computations based on the model described in section 4.3.
Traveling-wave systems 151

4.5 Experiments on a thermoacoustic refrigerator

4.5.1 Introduction
A thermoacoustic-Stirling refrigerator, which utilizes a traveling wave to realize
the Stirling cycle within the regenerator, was built to explore the performance at
high frequency and relatively small size. A further objective of this experimental
set-up was validation of the model described in section 4.3. Helium was chosen as
the working gas because of its suitable properties to produce more cooling power
than the other gases under the same working conditions. However, the sound speed
in helium is much larger than in the other gases leading to larger size at the same
operation frequency. The refrigerator is driven by a pressure wave generator (PWG)
provided by CFIC Q-Drive.
The design considerations are discussed in more detail in the next section. After
that, in section 4.5.3, the resulting experimental set-up is described and section
4.5.4 focuses on the measuring equipment and data handling. Losses in the set-up
are discussed in section 4.5.5. The measuring procedure is considered in section
4.5.6 followed by measuring results and discussion in section 4.5.7. This section
ends with conclusions in 4.5.8.

4.5.2 Design considerations


1. The diameters of tube sections were chosen from commercially avalaible
products. The coupling in size between tubes and tee was also taken into
account. With all these considerations, the tubes in the loop and tee were
chosen to be of stainless steel inner diameter 29.7 mm and 2 mm wall thickness.
2. The operation frequency is determinded by the operation frequency of the
PWG provided by CFIC Q-Drive. Originally, two theoretical designs of the
set-up were made, one operating at 400 Hz and the other one at 1000 Hz.
However, the maximum that Q-Drive could provide us with was 320 Hz PWG.
Therefore, the design was adapted to 320 Hz.
3. The system was designed as a half-wave length system. Therefore, the
resonator tube length was roughly determined by the operation frequency of
320 Hz.
4. The rough length of the loop was determined by using the model of section 4.3,
and refined later by using a numerical simulation solver from the research
group of E. Luo and W. Dai [73]. This also determines the length of the
regenerator and the tube segment connecting the cold-end heat exchanger and
the tee. As stated before, the author found that a lumped volumetric
compliance is not necessary for a good performance. Therefore, in the design,
152 Chapter 4

there was no compliance. A single tube diameter also allows for simpler
manufacturing and assembly.
5. The position of regenerator was determined by using the above-mentioned
numerical simulation solver to place it at the position with a large acoustic
pressure and a small volume velocity position.
6. The regenerator material, dimensions of stainless steel wire screen were chosen
based on the operation frequency and the conclusive selection criteria in
section 4.4.
7. All the dimensions were refined by using the numerical simulation solver
mentioned under point 4.

4.5.3 Experimental set-up


The structure of the system and its 3-D schematic drawing are given in Figs. 4.5.1 a
and b, respectively. Essentially, it consists of a half-wavelength resonator filled
with helium, with one end connected to the pressure wave generator and the other
connected to a torus-shaped section. The torus-shaped section contains three heat
exchangers, a regenerator and a duct forming a loop. Six microphones and five
thermometers are mounted along the tubes and connected to the data-acquisition
equipement. A photograph of the complete system is given in Fig.4.5.1 c.

Ambient heat exchanger


Regenerator
Pressure wave Electrical heater
generator Thermal buffer
Secondary heat exchanger

Figure 4.5.1(a): Structure of the thermoacoustic-Stirling refrigerator.


Traveling-wave systems 153
Pressure wave generator
P8
P7

Microphones

Average pressure
sensor P9

Figure 4.5.1(b): Experimental set-up of the thermoacoustic-Stirling refrigerator


(Two separate tube sections are included for tuning the resonator length).

Pressure wave
generator
Resonator tube

The loop
Figure 4.5.1(c): Photo of the thermoacoustic-Stirling refrigerator.

Pressure wave generator (PWG)


The acoustic power is generated by a powerful pressure wave generator, model
2s102W-X by Q-Drive. It has a piston area of 91.5 cm² and motor stiffness of
34.3kN/m. The wave generator is fixed by bolts and nuts on an aluminium profile.
Two microphones (P7 and P8) are installed in two chambers to monitor the
acoustic pressure inside the PWG.

The resonator tube


The resonator tube is made of a stainless steel tube with uniform inner diameter of
38.4 mm and wall thickness of 2 mm. It consists of three segments (1, 2, and 3) and
the total length is 1.2 m as designed. A schematic drawing is given by Fig.4.5.2.
Segment 1, which is 800 mm long, is surrounded by a larger diameter tube of 44.3
mm inner diameter and wall thickness of 2 mm. Water cooling circulation is
154 Chapter 4

applied in the gap between the two tubes. By doing so, the heat due to acoustic
dissipation along the wall can be removed and the amount of heat can be
determined by measuring the water temperatures in and out. Segment 2, which is
200 mm long, can be replaced by another tube of different length or their
combination to make different resonator lengths. The two optional tubes are 100
mm and 50 mm long. The segment tube 3 is 200 mm long. Along segment 3, two
microphones are installed to measure the acoustic power passing through, and one
static pressure sensor is for measuring the filling pressure inside the system. One
end of the resonator tube is coupled to the PWG by a flange and the other end is
connected to a reduced tube by a flange. The reduced tube has diameters of 38.4
mm and 29.7 mm, length of 26 mm.

Loop
section

Jacket water cooling

Segment 1 Segment 2 Segment 3 Reduced


PWG tube

Figure 4.5.2: Schematic drawing of the thermoacoustic-Stirling refrigerator.

The loop
The reduced tube is connected to one branch of the tee junction. The junction is a
commercial available, stainless-steel tee with inner diameter of 29.7 mm. One
branch of the tee is connected to the feedback tube, which is a torus tube section of
261 mm long with uniform inner diameter of 29.7 mm. One end of the feedback
tube is bent to connect to the tee and the other end connects to the tube segment
that contains the regenerator and the heat exchangers. The regenerator and the
thermal buffer tube are surrounded by a vacuum chamber filled with super
insulation to be thermally isolated from the environment (see also Fig.4.5.5 (a)).
Along the bent tube, three microphones are installed and a fourth microphone is
installed in the tee junction. The ambient heat exchanger is 6 mm thick. A
diaphragm is installed close to the ambient heat exchanger to prevent DC flow. The
photo of the ambient heat exchanger is given by Fig. 4.5.3.
The ambient heat exchanger was a delicate manufacturing product, involving much
precision machining work using spark discharging technique in copper. The proper
functioning of the ambient heat exchanger plays a key role in the performance of
the whole system. The high working frequency requests thin fins and a small
Traveling-wave systems 155

distance between any two fins. The thickness of a fin is 0.11 mm and the distance
between two adjacent fins is 0.17 mm, as shown by the photo of a prepared test
plate in Fig. 4.5.3. Keolian and Hofler have conducted a series of studies on heat
exchangers [74-77]. In the work of 1994 [74], Hofler concluded that a heat
exchanger with ξhx/Lhx in the range of 3 to 8 can be thermally effective as a source
or sink for thermoacoustic heat transport if y0/δκ is in the range of 0.75 to 0.5. Here,
the peak displacement amplitude of a gas parcel in the heat exchanger is given by
ξhx, and the length of the heat exchanger is given by Lhx. The separation between
adjacent fins is 2y0 and δκ is the thermal penetration depth of the working fluid.
The conclusion from Hofler was employed in the design of the heat exchangers in
this setup. The details of designing the ambient heat exchanger are given in
appendix I.

Figure 4.5.3: Photo of the ambient heat exchanger.

Figure 4.5.4: Photo of the electrical heater at the cold end of the regenerator.
156 Chapter 4

Diaphragm 29.7 mm
29.7 mm
Ambient heat exchanger 6 mm
Regenerator
Electrical heater Feedback 23 mm
tube
Thermal buffer chamber 4 mm
Secondary heat exchanger 29.7 mm
25 mm
Vacuum chamber 29.7 mm
6 mm
Tee

Figure 4.5.5(a): Schematic drawing of the loop section of the thermoacoustic-


Stirling refrigerator. The dimensions of the components inside the vacuum
chamber are given in the right-side drawing.

Tee

Vacuum
chamber

Feedback tube

Figure 4.5.5(b): Photo of the loop section of the thermoacoustic-Stirling


refrigerator.

The regenerator has a thickness of 23 mm and diameter of 29.7 mm. The


regenerator is made of a stack of stainless steel wire screens of Mesh 180 and wire
diameter of 0.039-0.04 mm. The regenerator is sandwiched between an electrical
Traveling-wave systems 157

heater and the ambient heat exchanger. In between two perforated plates are used
to fix the regenerator screens. The electrical heater forms a heat load to the
refrigerator and represents the cooling load. The heater is depicted in Fig. 4.5.4,
with the fin thickness of 0.11 mm and interfin distance of 0.17 mm. The thermal
buffer tube is 25 mm long, and is lined with a thermal isolation ring made of Peek.
One end of the thermal buffer tube is connected to a secondary heat exchanger of 6
mm thick. The second heat exchanger has the fin thickness of 0.12 mm and interfin
distance of 0.27 mm. The dimensions are given in Fig. 4.5.5 (a) and a photograph
of the complete loop is given in Fig. 4.5.5 (b).

4.5.4 Measuring equipment and data handling


Thermometers are mounted on the heat exchangers as well as in the water channels
of the cooling water. Pressure sensors are mounted flush with the wall at various
parts of the system. In this way the acoustic signal information as well as the heat
fluxes can be determined. Fig. 4.5.6 shows the distribution of all measuring devices.
A HP 33120A function generator generates a sinus wave that is used as input for a
Dynacord L2400 amplifier in bridged operation. The frequency and amplitude of
the sinus signal can be read from the control panel.

Julabo Grundfos
Flow
cooler pump P4
meter
T6 T5
T7
P5
V1 Q V2 Heater V, I T4
P7 T3
T9 P6
T11 T10 T8
P9 T1 T2
DC pressure
P2 P3 P1
Function Amplifier HIOKI PWG
P8
Generator power
sinus sensor P7, P8 via Endevco

EG&G Lockin

Figure 4.5.6: Overview of the measuring devices.

The amplified sinus signal passes through the Hioki power sensor and goes to the
PWG (model 2s102W-X by Q-Drive). Two microphones Endevco (P7, P8) are
mounted on the walls of the two back volumes of the pistons to monitor the
acoustic pressure inside the two chambers. The two microphones, P7 and P8, are
measured by an EG&G lockin type amplifier. Microphones P1 to P6, PCB
Piezotronics are mounted along the resonator tube and the loop. P9 is to measure
158 Chapter 4

the static mean pressure inside of the system. When the heater is on, the power is
provided by a power supply type E030-5 of Delta Elektronika. T1 to T5 are Pt1000
temperature sensors that are used to monitor the inside temperatures of the heat
exchangers. T2 is attached to the solid part of the secondary heat exchanger for
measuring the temperature of the metal material. T1 is attached to a support, which
is attached to the secondary heat exchanger, for measuring the oscillating gas
temperature flowing back and forth. In a similar way, T4 is to measure the
temperature of the solid part of the cold-end heat exchanger and T3 is to measure
the temperature of the gas. T5 is for measuring the temperature of the gas
oscillating through the ambient heat exchanger. T6 to T11 are to measure the
temperature of the water circulating inside the heat exchangers. Water is pumped
by a Grundfos pump and the temperature of the circulating water is maintained at
25°C by a Julabo F32 cooler. The water flow rate is measured by flowmeters
installed downstream of the Grundfos pump. T6 is the temperature of water
flowing in and T7 is the temperature of water flowing out of the ambient heat
exchanger. T8 and T9 are the temperatures of water flowing in and out of the
secondary heat exchanger. T10 and T11 are the in and out flow temperatures of the
water cooling pipes that cool the resonator tube wall. The thermometers T6-T11
have an accuracy of 0.01 K and are very relevant as these are needed to determine
the thermal fluxes of the apparatus. There is no displacement sensor (LVDT)
mounted on the piston of PWG. And due to time problem, the acoustic power at the
PWG was not measured in this experiment.
During a measurement, all data (except P7 and P8) are collected with a PXI system
by National Instruments and handled by a LabView program. The parameters are
shown on the LabView window and a text file can be generated for every
measuring point. The data written in the file can be read in a Matlab-file to
compute the desired quantities, which are discussed below.

Energy exchanged in heat exchangers


The heat released by oscillating gas through the ambient heat exchanger is
absorbed by the water circulation. The absorbed heat PambHX can be obtained by
PambHX = V&water ⋅ ρ water ⋅ c p water ⋅ (T 7 − T 6) , (4.5.1)
where V&water is the volume flow rate of the water through the heat exchanger.
Similarly, the energy exchange taking place at the secondary heat exchanger and
the water cooling of the resonator tube can be obtained by
Psec HX = V&water ⋅ ρ water ⋅ c p water ⋅ (T 9 − T 8) , (4.5.2)
PresHX = V&water ⋅ ρ water ⋅ c p water ⋅ (T 11 − T 10) . (4.5.3)
Cooling power
Traveling-wave systems 159

The cooling power at the cold end of the regenerator is dissipated by an electrical
heater. Therefore, the cooling power is given by
Pcooling = U ⋅ I , (4.5.4)
where U and I are the voltage and current of the power supply, respectively.
The microphones were mounted to determine the acoustic power flowing through
by the “two-microphone method” [62], but the results could not comply with
analysis of energy balance. In this configuration, phase differences were so small
that errors in the microphone readings (e.g. resulting from vibrations) deteriorated
the overall accuracy to such an extent that they could not be used for reliable
evaluations of efficiency. Therefore, the acoustic power computed from parameters
of these microphones is not used in this work.

Acoustic power
The acoustic power is obtained by analyzing the energy balance of the system. This
is shown by the control volume in the dashed rectangle in Fig. 4.5.7. Acoustic
power enters the control volume to drive the refrigerator. The control volume has
energy exchange with the external heat sources or sinks at the ambient heat
exchanger, secondary heat exchanger and the electrical heater at the cold-end heat
exchanger.

Penvironment

PambHX
Pcooling
PsecHX
Pacoustic

Figure 4.5.7: Energy balance of the loop.

The regenerator and thermal buffer tube are thermally insulated by a vacuum
chamber (see Fig.4.5.5 (a)). Therefore, the heat exchange of this part with the
environment is considered to be zero. Besides the heat exchange mentioned above,
the loop exchanges heat with the environment along the torus tube, as denoted as
Penvironment in Fig. 4.5.7. Considering the practical situation of the operation of the
setup, Penvironment is small enough to be neglected in energy balance analysis. In the
steady state, the energy conservation of the control volume gives
160 Chapter 4

Pacoustic + Pcooling = PambHX + Psec HX . (4.5.5)


The acoustic power flowing into the control volume can be given by
Pacoustic = PambHX + Psec HX − Pcooling . (4.5.6)

4.5.5 Losses
There are many possible losses in this system. Two of these can be derived from
the measurement results and will be extensively discussed here.

1. Viscous loss along the tube wall


The viscous loss converts part of the acoustic power into heat, which is absorbed
by the solid tube wall and will warm up the wall. Part of the resonator tube wall is
cooled down by water circulation between temperature measurement of T10 and
T11 as shown in Fig. 4.5.6. Based on the heat taken away by the water circulation
PresHX and the surface area of the tube wall between T10 and T11, the total viscous
loss of the complete refrigerator along the wall, denoted as Ptotal wall loss, can be
approximated by:
Ptotal wall loss = ( PresHX / A10 −11 ) ⋅ Atotal (4.5.7)
A10-11 is the inner surface area of the tube wall between T10 and T11. Atotal is the
total inner surface area of the tube wall of the complete refrigerator.
Later in the thesis the energy balance as depicted in Fig.4.5.7 is considered with the
dashed box (system boundary starting at microphone P2, see Fig.4.5.6).
The viscous loss along the tube wall of the part downstream the location of
microphone 2 can be approximated by:
Pwall loss = ( PresHX / A10 −11 ) ⋅ AMic 2 . (4.5.8)
AMic2 is the inner surface area of the tube wall downstream microphone P2.

2. Convection streaming losses


When the driver is switched on, the temperature at the cold end of the regenerator
goes down rapidly. A temperature gradient builds up between the cold end of the
regenerator and the secondary heat exchanger. Flows inside the system cause
convective heat exchange, as shown in Fig. 4.5.8. Note that this convective heat
exchange is not only natural convection but may be caused by any streaming in the
system as well. The zone where the convection takes place is indicated. Since the
temperature at the secondary heat exchanger T2 is anchored at environmental
temperature, we indicate it as the thermally grounded case. If the secondary heat
exchanger is thermally floating, the temperature T2 will fall with the temperature
T4 of the cold end of the regenerator. In that case, T2 is lower than the
environmental temperature and higher than T4.
Traveling-wave systems 161

T4

Pconvection Ambient HX

T2
Secondary HX

P2

Figure 4.5.8: Convection loss in the thermal buffer tube in thermal grounded case
caused by streaming.

T4
Ambient HX
T2
Pconvection

T2 Secondary HX

P2

Figure 4.5.9: Convection loss in the thermal buffer tube in thermal floating case.

Ambient Secondary heat


environment P02 exchanger P24 Regenerator

R02 R24 T5

T0 T2 T4

Figure 4.5.10: Thermal resistance model for convection loss in the loop part of the
refrigerator system.

Due to the temperature gradients between T2 and T4, and between T2 and its
surrounding environment, heat exchange by convection arises, as shown in Fig.
4.5.9. As an assumption, the convective heat exchange is proportional to the
temperature difference between two surfaces, using Newton’s law of heat transfer
Q& = α ⋅ (TA − TB ) . Therefore, the convection loss can be modeled and computed
for the thermally floating case, in which it is not measured.
162 Chapter 4

Assume that the cold end of the regenerator with temperature T4 is connected with
the secondary heat exchanger by a thermal resistance R24, and that the secondary
heat exchanger is connected with the environment with temperature T0 by thermal
resistance R02, as shown in Fig. 4.5.10.
In Fig. 4.5.10, the energy flow between the environment and the secondary heat
exchanger is P02. Similarly, the energy flow between the secondary heat exchanger
and the cold end of the regenerator is P24. Concerning the secondary heat exchanger,
there are two cases given below.
(A) Thermally grounded condition
When the secondary heat exchanger is maintained at environmental temperature by
water circulation through it, there is no heat exchange between the environment
and the secondary heat exchanger. That means the energy exchange shown in the
dashed oval in Fig. 4.5.9 does not take place, i.e. P02=0. The energy flow between
the secondary heat exchanger and the cold end of the regenerator P24 is due to
internal convection and is balanced by water circulation through the secondary heat
exchanger PsecHX. Therefore, it can be written as:
⋅ (T 4 − T 2 ) .
1
Psec HX = P24 = (4.5.9)
R24
By using the corresponding data from measurement in the thermal grounded case,
the thermal conductance 1/R24 can be obtained, which is then considered to be
constant.
(B) Thermally floating condition
When the secondary heat exchanger is floating, the temperature T2 of the
secondary heat exchanger will drop when the cold end of the regenerator T4 is
cooling down. There is no thermal insulation applied at the tee junction part, the
energy flow P02 takes place as shown in the dashed oval in Fig. 4.5.9. For this case
holds P02=P24. The convection loss can be obtained by:
⋅ (T 4 − T 2 ) ,
1
Pconvection = P24 = (4.5.10)
R24
where 1/R24 was obtained from the thermally grounded case as discussed above.

4.5.6 Measuring procedure


Filling the system
Every time when the system is opened, the system has to be refilled with pure
helium gas before conducting a new measurement. To refill it by pure helium gas,
the system was flushed three times. During filling and releasing gas, the mean
pressure (filling pressure) is monitored by the static pressure sensor P9.
Two kinds of measurement were carried out, indicated as operation A and B:
Traveling-wave systems 163

Operation A: Maximum temperature difference across the regenerator (zero


cooling load)
This is also the measurement of the lowest temperature that the cold end of the
regenerator can reach. In this case, the electrical heater at the cold end of the
regenerator was off, so no added cooling load on the cold end. The PWG is set at
10 W, which can be read off from the Hioki power meter. It generates direct input
to the PWG, and this fixed value is maintained. After thirty minutes of operation,
the steady state has established. The data were stored in file. Then the input power
was stepwise increased to 30, 50 and 70 W. At each input power, the same
operation was repeated.
Operation B: Performance as a function of cooling power
In this measurement, the electrical heater at the cold end of the regenerator was
powered to generate a heat load. Measurements were made at fixed inputs (30 W,
50 W, 70 W) to the PWG. At every input, the heater power was varied, and the
system parameters were stored in file as steady state was reached.
All the data files are processed using a Matlab program. In every file, around 30
values are recorded for each parameter. Therefore, the mean value for each
parameter is calculated and used for related energy computation in the Matlab
program.

4.5.7 Results and discussion


In the experiments, two orientations of the loop section are investigated: upward
and downward with varied input acoustic power under thermally grounded
condition. The results showed that the performances of two orientations are close
but the downward configuration behaved more stable than the upward. The time
evolutions of the two orientations are given in appendix J. Therefore, all the
measurements presented later in this work were conducted with the loop section in
downward configuration. In the downward configuration, any unstability caused by
natural convection between the secondary heat exchanger and the cold end of the
regenerator is weakened.

I. Comparison of performance of different resonator tube lengths


In the design phase, the resonator tube length was optimized at filling pressures of
11 bar and 15 bar. The resonator tube introduced in section 4.5.3 is shown in Fig.
4.5.11. There are three microphones installed on segment 3. So neither segment 3
nor segment 1 can be removed from the system. Only segment 2 can be replaced by
another tube of different length or their combinations. There are four options: (a)
200 mm, (b) 100 mm, (c) 50 mm and (d) no tube (0 mm). The length containing
only tube a (200mm) as segment 2 is the design value. The lengths larger and
164 Chapter 4

smaller than the designed length were measured by conducting operation A in


thermally grounded condition. The measured combination of the optional tubes is
listed in Table 4.5.I. Some cases did not show resonance and are indicated with
“No” in Table 4.5.I.

Segment 1 Segment 2 Segment 3

800 mm
200 mm

Figure 4.5.11: Three segments of the resonator tube.

Resonant Tube a Tube b Tube c Tube d Resonator


frequency 200 mm 100 mm 50 mm 0 mm length (mm)
Yes √ √ √ × 1350
No √ × √ × 1250
Yes √ × × × 1200
Yes × √ × × 1100
No × × × √ 1000
Table 4.5.I: Measured resonator tube segment 2 of optional tubes and their
combinations. “√” indicates “used”, and “×” indicates “not used”.

The cases without a resonant frequency are not discussed here. The comparison of
three lengths of resonator tube at filling pressure of 11 bar and 15 bar are given in
Fig. 4.5.12 and 4.5.13. ∆T is the temperature difference across the regenerator, i.e.
∆T=T5-T4.
From the comparison in Figs. 4.5.12 and 4.5.13, the designed length (1200 mm)
has the best performance. In section 4.3, the strong dependency of the complete
acoustic impedance of the refrigerator on the resonator tube length is shown in Figs.
4.3.8 and 4.3.9. Subsequently, the complete acoustic impedance of the refrigerator
has strong influence on the final performance through the coupling between the
refrigerator and the PWG. It can be understood that the resonance of the complete
system (refrigerator and the PWG) only takes place when the impedances of the
refrigerator and the PWG are matched (reactance part of the impedance of the
complete system, refrigerator and the PWG, is zero). In the design phase, the
length of the resonator tube was optimized by numerical computation at the
designed working condition. In Fig. 4.5.13, the performance of 1350 mm is close to
that of the designed length 1200 mm. Because Pmean=15 bar is not the designed
Traveling-wave systems 165

working condition. The length of 1200 mm might not be the optimum value in that
specific case.

45

40 1200 mm
1350 mm
35 1100 mm
Pmean=11 bar
30

25
∆ T (K)

20

15

10

0
0 10 20 30 40 50 60 70
Pdriver (W)

Figure 4.5.12: Temperature difference across the regenerator for three different
lengths of resonator tube at filling pressure of 11 bar.

40
1200 mm
1350 mm
35 1100 mm
Pmean=15 bar
30

25
∆ T (K)

20

15

10

0
0 10 20 30 40 50 60 70
Pdriver (W)

Figure 4.5.13: Temperature difference across the regenerator for three different
lengths of resonator tube at filling pressure of 15 bar.
166 Chapter 4

II. Comparison of performance of thermally grounded and floating cases on the


maximum temperature difference across the regenerator with zero cooling load
In order to investigate the temperature characteristics of the apparatus with the
resonator at its designed length, the so called zero-load measurements (operation A)
were made. In that case, no power is supplied to the heater and the thermoacoustic
cooler reaches its lowest temperature. Zero-load measurements (operation A) were
made for thermally grounded and floating cases at filling pressures of 11 bar, 12
bar and 15 bar. The results are given in Fig. 4.5.14.
The thermally floating cases show a better performance than the thermally
grounded ones. The temperature difference ∆T in the thermally grounded cases is
smaller due to the heat exchange between cold-end heat exchanger and the
secondary heat exchanger as discussed in section 4.5.5. In the thermally floating
cases, the secondary heat exchanger cools down along with the cold-end heat
exchanger, leading to less heat loss from T2 to T4, and thus a larger ∆T can result.
Computation by using the model described in section 4.3 and corresponding
measurement of thermally floating cases is given in Fig. 4.5.15. In the model, there
is no secondary heat exchanger. Therefore, the thermally floating case is the closest
to the situation in the model.

Pmean=11 bar (thermally grounded)


Pmean=12 bar (thermally grounded)
60 Pmean=15 bar (thermally grounded)
Pmean=11 bar (thermally floating)
55
Pmean=12 bar (thermally floating)
50 Pmean=15 bar (thermally floating)
45

40

35
∆T (K)

30

25

20

15

10

0
0 10 20 30 40 50 60 70
Pdriver (W)

Figure 4.5.14: Zero-load temperature difference across the regenerator in


thermally grounded and floating cases at various filling pressure.
Traveling-wave systems 167

In all computations, the temperature at the hot end of the regenerator, TH, is
assumed to be at ambient temperature. The filling pressure (mean pressure Pm) is
set at a constant value corresponding with the measurement. The operating
frequency is set at the corresponding measured value. The working gas is helium
and the regenerator material is stainless steel wire screens. The properties of the
working gas and the regenerator material, such as thermal conductivity, ratio of
specific heat, Prandtl number, used to compute Eqs. (4.3.37) and (4.3.52) can be
found in Refs. [38,39]. The geometrical parameters (L-fb d-fb L-cpl d-cpl L-reg d-reg rh
ψreg L-tb d-tb L-res d-res) are given in section 4.5.3. The acoustic pressure measured at
P2, as shown in Fig. 4.5.6, is given as input acoustic pressure p1in and increased
stepwise. The distance between microphone P2 and the center of the tee junction is
set at Lres. For every p1in, an iteration was executed to obtain the lowest temperature
for this p1in. In one iteration, the cold-end temperature of the regenerator TC was
decreasingly varied in a range of values. Therefore, for one calculation in an
iteration, a specific p1in and TC are inputed into Eq. (4.3.52) and related equations,
and the cooling power can be obtained. In the next calculation in the same iteration,
a different TC and the same p1in are inputed into Eq. (4.3.52) and related equations,
and the cooling power for this TC and p1in can be obtained. After calculations for
many TC and the same p1in in one iteration, many cooling power values are
obtained. Since TC is given in a decreasing manner, the cooling power values are
obtained in a decreasing manner as well. When the cooling power is almost zero,
the cold-end temperature of the regenerator TC is stored as the lowest temperature
Tlowest at this working condition. Thus the temperature difference across the
regenerator ∆T (∆T=TH –Tlowest) at this input acoustic pressure p1in is obtained.
Then another iteration for the next p1in is repeated.
In Fig. 4.5.15, the trends from computation and the corresponding measurement are
the same. The computation results overpredict the experimental values with a
factor of about 2. The discrepancy between computation and experimental
measurement can be explained by the highly idealized assumptions made to build
the analytical model. All the losses, except dissipation and heat conduction within
the regenerator, are neglected in the model. Some of these losses are analyzed in
more detail below.
Both the computation and the measuremts show that the performance decreases
with increasing filling pressure Pmean. This can be explained by the fact that the
configuration is optimized at a filling pressure of 11 bar in the design. When Pmean
becomes larger, the gas thermal penetration depth δκ becomes smaller. With the
same regenerator material, the ratio of hydraulic radius rh to thermal penetration
depth δκ becomes larger. This means that the heat exchange becomes worse.
168 Chapter 4

computation Pmean=11 bar


experiment Pmean=11 bar
computation Pmean=12 bar
experiment Pmean=12 bar
100 computation Pmean=15 bar
90
experiment Pmean=15 bar

80

70

60
∆T (K)

50

40

30

20

10

0
4 4 4 4 4
1.0x10 1.5x10 2.0x10 2.5x10 3.0x10
P2 (Pa)

Figure 4.5.15: Temperature difference across the regenerator of thermal floating


cases and computation at various filling pressure.

Some energy flows and losses


As stated above, losses play an important role in the performance of the system.
From the many possible loss mechanisms, we limit the following discussion to the
losses considered in section 4.5.5.
In addition to the losses, some energy flows in this set-up, which are important for
understanding the energy balance of the whole system, are presented. Comparison
between the energy flows and the losses can help to understand the discrepancy
between computation results and the measurements. Both thermally grounded and
floating cases are considered for further comparison.
The heat, exchanged at the ambient heat exchanger PambHX , is obtained by using Eq.
(4.5.1). The heat exchange at the secondary heat exchanger PsecHX is given by Eq.
(4.5.2) in the thermally grounded case, and is zero in the thermally floating case. In
the measurement of the largest attainable temperature difference across the
regenerator (operation A), the electrical heater is off. Therefore the cooling power
is zero in using Eq. (4.5.6) to obtain Pacoustic.
The viscous losses along the tube wall Ptotal wall loss and Pwall loss are given by Eqs.
(4.5.7) and (4.5.8). By using the corresponding data (PsecHX, T4 and T2) from
measurements in the thermally grounded condition and Eq. (4.5.9), the thermal
Traveling-wave systems 169

conductance 1/R24 can be obtained by linear fit of PsecHX and temperature difference
between T4 and T2 by linear fit tool in software Origin. Eq.(4.5.9) shows that the
linear relationship between PsecHX and the difference of T4 and T2 (T4-T2) has to
pass the original point (0,0) in the linear fit curve. The linear fit of case Pmean=11
bar is taken as an example as shown in Fig. 4.5.16. It seems that there exists a
better linear fit, which will pass through more measurement points than the line
illustrated in Fig. 4.5.16. But that line would not pass through the original point (0,
0). This method of linear fit was repeated for the other two cases of Pmean=12 and
15 bar. The resulting conductance 1/R24 for these filling pressures is listed in the
table 4.5.II. The convection loss in the thermally floating cases is obtained by using
Eq. (4.5.10) and corresponding value of thermal conductance 1/R24 in table 4.5.II.

Pmean 11 bar 12 bar 15 bar


1/R24 (W/K) 0.366 0.321 0.409
Table 4.5.II: Conductance 1/R24 derived from Eq.(4.5.9) at various filling
pressures.

-2

measured value
-4 linear fit (by Eq.(4.5.9))

-6
PsecHX (W)

-8

-10

-12

-14
-32 -30 -28 -26 -24 -22 -20 -18 -16 -14 -12
T4-T2 (K)

Figure 4.5.16: Linear fit of temperature difference T4-T2 and PsecHX by Origin for
the case of Pmean=11 bar in thermally grounded condition.

The energy flows and losses for a filling pressure of 11 bar are given in Figs.
4.5.17 to 4.5.19. The Figs. 4.5.20 to 4.5.22 are for a filling pressure of 12 bar and
from 4.5.23 to 4.5.25 are for a filling pressure of 15 bar. Figs. 4.5.17, 4.5.20, and
4.5.23 are plots for energy flows and viscous losses with respect to the complete
170 Chapter 4

refrigerator. Since these were zero-load measurements (operation A), the cooling
power is zero, i.e. Pcooling=0. For the thermally grounded cases, the energy flows
taking place at the ambient heat exchanger and the secondary heat exchanger, and
the total wall loss defined in Eq. (4.5.7) for viscous loss of the complete inner
surface of the refrigerator are shown in the figures. For the thermally floating cases,
the heat exchange at the secondary heat exchanger is zero. Therefore, only the
energy flows at the ambient heat exchanger and the total wall losses are shown in
the figures. Energy flowing from a heat exchanger to the outside world is defined
positive.
The energy flows at the secondary heat exchanger PsecHX in the three figures are
negative, because heat flows from the outside to that heat exchanger (since that is
cooled along with the cold heat exchanger).
Figs. 4.5.18, 4.5.21, and 4.5.24 are plots for energy flows and viscous losses in the
part downstream of the microphone 2 (P2 in Figs.4.5.6 and 4.5.9), which is the part
in the dashed rectangle in Fig. 4.5.7. The acoustic power going into the loop
section Pacoustic is calculated by using Eq. (4.5.6) and the viscous loss Pwall loss is
computed by using Eq. (4.5.8). The convection loss Pconvection is obtained by using
Eq. (4.5.10). In these three figures, the values of the viscous loss Pwall loss for
thermally grounded case are close to their corresponding points in thermally
floating case. The convection loss Pconvection is negative because of the thermal
coupling between cold heat exchanger and secondary heat exchanger (R24 in
Fig.4.5.10). The convection loss in the thermally grounded case is not separately
depicted since it equals the heat flowing to the secondary heat exchanger (PsecHX).
The acoustic power Pacoustic in the thermally floating cases is larger than the
corresponding one in thermally grounded cases. In the latter cases, more acoustic
energy is needed to compensate for the loss from the secondary heat exchanger to
the environment, which is shown by P02 in Fig. 4.5.10.
Figs. 4.5.19, 4.5.22 and 4.5.25 show that the sum of the losses (Pconvection and Pwall
loss) is comparable to the input acoustic power. Therefore, neglecting these losses in
the analytical model of section 4.3 will be the main cause of the discrepancy
between computation results and measurements as shown in Fig. 4.5.15.
Furthermore, note that the losses are about half the acoustic power, these also
explaning the factor of two difference between computation and measurements.
Traveling-wave systems 171

PambHX (thermally grounded)


PsecHX (thermally grounded)
Ptotal wall loss (thermally grounded)
30
PambHX (thermally floating)
25 Ptotal wall loss (thermally floating)
Pmean=11 bar
20

15
Power (W)

10

0
0 10 20 30 40 50 60 70
-5 Pdriver (W)

-10

-15

Figure 4.5.17: Energy flows and viscous loss along the wall of thermally grounded
and thermally floating cases at filling pressure of 11 bar.

Pacoustic (thermally grounded)


PsecHX (thermally grounded)
Pwall loss (thermally grounded)
Pacoustic (thermally floating)
22 Pconvection (thermally floating)
20
Pwall loss (thermally floating)
18
16 Pmean=11 bar
14
12
10
8
Power (W)

6
4
2
0
-2 0 10 20 30 40 50 60 70
-4
Pdriver (W)
-6
-8
-10
-12
-14

Figure 4.5.18: Energy flows and losses in the part downstream of microphone 2 of
thermally grounded and thermally floating cases at filling pressure of 11 bar.
172 Chapter 4

22
Pacoustic (thermally floating)
20
Pconvection+Pwall loss
18 (thermally floating)
Pmean=11 bar
16

14
Power (W)

12

10

0 10 20 30 40 50 60 70
Pdriver (W)

Figure 4.5.19: Input acoustic power at microphone 2 and sum of losses in the part
downstream of microphone 2 of thermally floating case at filling pressure of 11 bar.

PambHX (thermally grounded)


PsecHX (thermally grounded)
Ptotal wall loss (thermally grounded)
PambHX (thermally floating)
25 Ptotal wall loss (thermally floating)
Pmean=12 bar
20

15
Power (W)

10

0
0 10 20 30 40 50 60 70
-5 Pdriver (W)

-10

Figure 4.5.20: Energy flows and viscous loss along the wall of thermally grounded
and thermally floating cases at filling pressure of 12 bar.
Traveling-wave systems 173

Pacoustic (thermally grounded)


PsecHX (thermally grounded)
Pwall loss (thermally grounded)
Pacoustic (thermally floating)
24 Pconvection (thermally floating)
22 Pwall loss (thermally floating)
20 Pmean=12 bar
18
16
14
12
10
Power (W)

8
6
4
2
0
-2 0 10 20 30 40 50 60 70
-4 Pdriver (W)
-6
-8
-10
-12

Figure 4.5.21: Energy flows and losses in the part downstream of microphone 2 of
thermally grounded and thermally floating cases at filling pressure of 12 bar.

25
Pacoustic (thermally floating)
Pconvection+Pwall loss
20 (thermally floating)
Pmean=12 bar

15
Power (W)

10

0
0 10 20 30 40 50 60 70
Pdriver (W)

Figure 4.5.22: Input acoustic power at microphone 2 and sum of losses in the part
downstream of microphone 2 of thermally floating case at filling pressure of 12 bar.
174 Chapter 4

PambHX (thermally grounded)


PsecHX (thermally grounded)
Ptotal wall loss (thermally grounded)
35 PambHX (thermally floating)
30 Ptotal wall loss (thermally floating)
Pmean=15 bar
25

20
Power (W)

15

10

0
0 10 20 30 40 50 60 70
-5
Pdriver (W)
-10

-15

Figure 4.5.23: Energy flows and viscous loss along the wall of thermally grounded
and thermally floating cases at filling pressure of 15 bar.

Pacoustic (thermally grounded)


PsecHX (thermally grounded)
Pwall loss (thermally grounded)
Pacoustic (thermally floating)
Pconvection (thermally floating)
25 Pwall loss (thermally floating)
Pmean=15 bar
20

15
Power (W)

10

0
0 10 20 30 40 50 60 70
-5 Pdriver (W)

-10

-15

Figure 4.5.24: Energy flows and losses in the part downstream of microphone 2 of
thermally grounded and thermally floating cases at filling pressure of 15 bar.
Traveling-wave systems 175

30
Pacoustic (thermally floating)
Pconvection+Pwall loss
25 (thermally floating)
Pmean=15 bar

20

Power (W) 15

10

0
0 10 20 30 40 50 60 70

Pdriver (W)

Figure 4.5.25: Input acoustic power at microphone 2 and sum of losses in the part
downstream of microphone 2 of thermally floating case at filling pressure of 15 bar.

III. Comparison of performance at varying cooling power for thermal grounded and
floating cases
In this measurement, the electrical heater is switched on to provide a heat load at
the cold end of the regenerator (operation B). The cooling power is obtained by Eq.
(4.5.4). The input power into the PWG was fixed at 30, 50 and 70 W for every
measurement. The filling pressure Pmean was fixed at 11 and 15 bar.
The temperature difference across the regenerator at various cooling powers is
plotted for the thermally grounded, and the thermally floating cases. The COP,
given by Eq. (4.5.11), at various cooling power is also plotted for both cases.
COP = Pcooling / Pacoustic . (4.5.11)
The Carnot COP is the maximal theoretical performance that a refrigerator can
achieve, and it is given by:
TC
COPC = . (4.5.12)
TH − TC
The relative COP, the ratio of COP to Carnot COP, is calculated as:
COPR = COP / COPC . (4.5.13)
176 Chapter 4

40
38 Pdriver=30 W
36 Pdriver=50 W
34
Pdriver=70 W
32
Pmean=11 bar
30
thermally grounded
28
26
24
∆T (K)

22
20
18
16
14
12
10
8
6
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.26: Temperature difference across the regenerator as a function of


cooling power at filling pressure 11 bar for thermally grounded case.

38
Pdriver=30 W
36 Pdriver=50 W
34 Pdriver=70 W
32 Pmean=15 bar
30 thermally grounded
28
∆T (K)

26
24
22
20
18
16
14
12

0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.27: Temperature difference across the regenerator as a function of


cooling power at filling pressure 15 bar for thermally grounded case.
Traveling-wave systems 177

55
Pdriver=30 W
50 Pdriver=50 W
Pdriver=70 W
45 Pmean=11 bar
thermally floating
40

35
∆T (K)
30

25

20

15

10
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Cooling power (W)

Figure 4.5.28: Temperature difference across the regenerator as a function of


cooling power at filling pressure 11 bar for thermally floating case.

55
Pdriver=30 W
50 Pdriver=50 W
Pdriver=70 W
45 Pmean=15 bar
thermally floating
40
∆T (K)

35

30

25

20

15
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.29: Temperature difference across the regenerator as a function of


cooling power at filling pressure 15 bar for thermally floating case.

As shown in the Figs. 4.5.26 and 4.5.27 for thermally grounded, and Figs. 4.5.28
and 4.5.29 for thermally floating cases, for all total input powers, the ∆T decreases
nearly linearly with increasing cooling power. Therefore, ∆T as a function of
cooling power Pcooling can be given as:
178 Chapter 4

∆T P
= 1 − cooling . (4.5.14)
∆Tmax Pcooling max
Here, ∆Tmax is the maximum ∆T for a specific input power, which is the
temperature difference at zero cooling load. Pcooling max is the maximum cooling
power that the system achieves for a specific input power, which is obtained at
∆T=0.
At relatively small levels of cooling power, the acoustic input power is merely
required to compensate for the losses. Therefore, the required input power will
hardly change as the cooling power is increased. Thus it can be expected that at
small levels of cooling power, the COP is proportional to that cooling power:
COP ∝ Pcooling . (4.5.15)
This is confirmed by Figs. 4.5.30-4.5.33.

Pdriver=30 W
2.2 Pdriver=50 W
2.0 Pdriver=70 W
Pmean=11 bar
1.8
thermally grounded
1.6

1.4

1.2
COP

1.0

0.8

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.30: COP as a function of cooling power at filling pressure 11 bar for
thermally grounded case.
Traveling-wave systems 179

1.8 Pdriver=30 W
Pdriver=50 W
1.6 Pdriver=70 W
Pmean=15 bar
1.4
thermally grounded
1.2

COP 1.0

0.8

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.31: COP as a function of cooling power at filling pressure 15 bar for
thermally grounded case.

1.4

1.2

1.0

0.8
COP

0.6

Pdriver=30 W
0.4 Pdriver=50 W
Pdriver=70 W
0.2 Pmean=11 bar
thermally floating
0.0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Cooling power (W)

Figure 4.5.32: COP as a function of cooling power at filling pressure 11 bar for
thermally floating case.
180 Chapter 4

1.0
Pdriver=30 W
Pdriver=50 W
Pdriver=70 W
Pmean=15 bar
0.8
thermally floating

0.6
COP

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.33: COP as a function of cooling power at filling pressure 15 bar for
thermally floating case.

The relative COPs are given in Figs. 4.5.34 to 4.5.37 for both thermally grounded
and floating cases. In these figures, the maximum values for most of the curves are
not yet reached in the range of cooling powers which were tested.
By using Eq. (4.5.12), the Carnot COP can be written as:
TC T − (TH − TC ) TH
COPC = = H = −1. (4.5.16)
TH − TC TH − TC ∆T
By using Eqs. (4.5.15) and (4.5.16), the relative COP is given as
COP Pcooling ∆T Pcooling
COPR = ∝ = . (4.5.17)
COPC TH / ∆T − 1 TH − ∆T
By employing Eq. (4.5.14), Eq. (4.5.17) can be rewritten as:
 P 
Pcooling 1 − cooling 
 Pcooling max 
COPR ∝   .
(4.5.18)
TH  P 
− 1 − cooling 
∆Tmax  Pcooling max 
A maximum COPR is achieved in Eq.(4.5.18) if
2
Pcooling  T   T   T 
=  H − 1 +  H − 1 −  H − 1 (4.5.19)
Pcooling max  ∆Tmax   ∆Tmax   ∆Tmax 
In the measurements, TH>>∆Tmax, thus Eq.(4.5.19) reduces to:
Traveling-wave systems 181

1
Pcooling = Pcooling max . (4.5.20)
2
In our measurements data ∆Tmax was in all experiments smaller than 50 K.
Substituting TH=300 K in that case, Eq.(4.5.19) would yield a ratio of 0.48 instead
of 0.5 which is obtained by assuming TH/∆Tmax→∞. So, a maximum in the relative
efficiency is expected at a value of Pcooling that is around half Pcooling max. When
extrapolating the curves in Figs. 4.5.26, we find for the grounded case at 30 W,
Pcooling max=7.5 W; at 50 W Pcooling max=13.7 W; and at 70 W Pcooling max=17W.
Maxima in the relative COP could thus be expected at 3.8 W, 6.9 W and 8.5 W,
respectively. Fig. 4.5.34 shows that the actual maxima are at slightly higher values
of cooling power. This is due to the fact that the COP increases more than
proportional to the cooling power as shown in Fig. 4.5.30. The same analysis can
be given for the other cases.

6
Relative COP [%]

4 Pdriver=30 W
Pdriver=50 W
3
Pdriver=70 W
Pmean=11 bar
2
thermally grounded
1

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.34: Relative COP as a function of cooling power at filling pressure 11


bar for thermally grounded case.
182 Chapter 4

6
Relative COP [%]

5
Pdriver=30 W
4
Pdriver=50 W
3
Pdriver=70 W
Pmean=15 bar
2 thermally grounded

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.35: Relative COP as a function of cooling power at filling pressure 15


bar for thermally grounded case.

6
Relative COP [%]

5 Pdriver=30 W
Pdriver=50 W
4
Pdriver=70 W
3 Pmean=11 bar
thermally floating
2

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Cooling power (W)

Figure 4.5.36: Relative COP as a function of cooling power at filling pressure 11


bar for thermally floating case.
Traveling-wave systems 183

Relative COP [%]


5

4
Pdriver=30 W
3 Pdriver=50 W
Pdriver=70 W
2 Pmean=15 bar
thermally floating
1

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Cooling power (W)

Figure 4.5.37: Relative COP as a function of cooling power at filling pressure 15


bar for thermally floating case.

4.5.8 Discussion
Measurement data obtained from the experimental set-up described in section 4.5.3
were compared to computation results based on the analytical model described in
section 4.3. The model correctly predicts the trends but overestimates the cooling
performance by about a factor of 2. This is due to the fact that losses due to viscous
effects and convection arise in practice that are not included in the model.
Many other research groups found similar losses in thermoacoustic-Stirling type
refrigerators. Tijani and Spoelstra developed and measured a coaxial
thermoacoustic-Stirling cooler [78]. This cooler has a different feedback structure
from the toroidal one and thus the loop section is more compact. Argon at a mean
pressure of 15 bar is the working gas and it operates at 60 Hz. The cooler has
achieved a relative efficiency of 25% of Carnot and a low temperature of -54°C
without heat load. Tijani and Spoelstia also measured the acoustic power dissipated
in the resonator by thermo-viscous processes at the wall as a function of the drive
ratio. They found that the viscous losses in the resonator at the operating point of
high performance (drive ratio=4.3%) reached as high as about half the power input.
Therefore, they proposed to improve the design by using a quarter-wavelength
resonator instead of a half-wavelength resonator and thus reducing the resonator
losses.
184 Chapter 4

Luo and Dai have been working on the thermoacoustic-Stirling refrigerator with
toroidal structure and driven by a thermoacoustic-Stirling engine [73, 79, 80]. They
built a traveling-wave thermoacoustic refrigerator driven by a traveling-wave
thermoacoustic engine and carried out numerical simulations about this set-up [73].
Their set-up achieved a lowest temperature of -64.4 °C and 250 W cooling power
at -22.1 °C when the system was filled with 3.0 MPa helium gas, working at 57.7
Hz and 2.2 kW of heat input into the thermoacoustic engine. They investigated the
performance at different heating powers and mean pressures. They found that their
numerical simulation program worked well in the frequency prediction and
reasonably well in the prediction of refrigerator inlet pressure amplitudes. But they
saw a large discrepancy between computation and measurement in the prediction
of the refrigerator’s cooling power as a function of cold-end temperature. They
attributed this large discrepancy to the serious underestimation of some loss
mechanisms strongly related to the thermoacoustic refrigerator’s cold-end
temperature, such as Rayleigh streamings inside the refrigerator’s thermal buffer
tube. This loss has similarity with the loss indicated as Pconvection in section 4.5. In
Figs.4.5.18, 4.5.21 and 4.5.24, the loss Pconvection is much larger than the viscous loss
along the tube wall Pwall loss and it is comparable to the acoustic power Pacoustic in the
thermally floating cases. Luo and Dai explained that neglecting the influence of
nonlinear behavior, especially the turbulence, in the model could be another cause
of the large discrepancy. This has similarities with the presented comparison
between computation and experimental measurements as shown in Fig.4.5.15. In
the work in [80], Luo and Dai theoretically investigated a thermoacoustically
driven Stirling cryocooler working at a high frequency of 500 Hz. By numerical
simulations, they concluded that the high operating frequency significantly
decreases the refrigeration efficiency. The relative efficiency (COP/Carnot
efficiency) of this 500 Hz refrigerator was predicted theoretically as about 8%-15%.
As discussed in the following chapter 5, the efficiency of a scaled-down system,
which will work at high frequencies as a result of scaling-down, decreases with
scaling factor.
Ueda et al. experimentally investigated pressure p and velocity U in the acoustic
field of a thermoacoustic Stirling engine with pressure sensors and a laser Doppler
velocimeter for flow visualization [81, 82]. They found that the negative phase lead
(around -20°) of U relative to p, rather than a pure traveling-wave phase (0°), made
the engine achieve a high efficiency. They inserted a second regenerator into the
looped tube to work as a cooler. Thus, the cooler used the work flow generated by
the engine to produce cooling [81, 83]. The cooler could reach a low temperature
of -25°C without heat load. They measured the lowest temperature at the cold end
TC of the cooler regenerator under different filling pressures when the total input
Traveling-wave systems 185

power into the hot end of the engine regenerator QH remained constant at 210 W.
The measurement showed an optimum filling pressure for each kind of working
gas. The cold-end temperature TC went higher when the filling pressure Pm was
larger or smaller than the optimum value. This has clear similarity to the trend in
Fig.4.5.15, where the filling pressure of 11 bar has the best performance in
comparison with 12 and 15 bar.

4.5.9 Conclusion
The analytical model, developed in section 4.3, is able to correctly predict the
trends, but overestimates the cooling performance by comparison with the
measurements results in section 4.5. The discrepancy is due to the highly idealized
assumptions in the model which exclude many losses taking place in practice. This
analytical model is acceptable for the next scaling analysis.
Chapter 5

Scaling considerations

5.1 Introduction

Many large-scale refrigerators were developed in the past decades. Small-scale


refrigerators also attracted some researchers’ attention. Many of these
investigations can be found in patents. In 2002, Chen reported on two miniature
standing-wave type refrigerators [48]. The maximum temperature difference was
10°C using an in-house MEMS stack material, under the operation conditions of 1
atm at around 4 kHz and 10 kHz. Symko has been devoted to the exploration of
miniature thermoacoustic refrigerators [49-51]. His miniature standing-wave type
coolers can be as small as a few centimeters long, and operating at thousands of
hertz. These impressively small thermoacoustic coolers have promising
applications on various industrial and electronics utilization.
Olson and Swift investigated similitude in thermoacoustics [86]. They proposed a
group of dimensionless parameters and rewrote the basic thermoacoustic equations
in dimensionless form. These dimensionless parameters are helpful to predict the
performance of a thermoacoustic device by measuring a scale model which
satisfies the similitude principles.
Benavides analyzed the minimum reachable size for thermoacoustic devices related
to the working gas [52]. He used the properties of the working gas to find the
limitation for scaling-down. In conclusion, he found that the reachable minimum
size for different gases could be the order of micrometer.
This chapter focuses on what during scaling-down could be the major problem that
limits miniaturization from the point of geometrical configuration and stack-
material construction. By using individual scaling factor in every dimension, the
performance can be predicted when a reference system is designed to shrink in any
or any combination of the three dimensions: x, y, z. Only geometry change is
considered, therefore, the changes don’t have to satisfy the similitude principles.
Scaling considerations 187

5.2 Standing-wave systems


In this chapter, the scaling of standing-wave systems is analyzed. It consists of
three subsections, referring to three different boundary conditions: constant
temperature difference over the stack, constant total energy flow and constant total
energy flow density.

5.2.1 Constant temperature difference over the stack


In this section, it is supposed that the temperature difference across the stack
remains unchanged while scaling down. We are going to focus on the variation of
the cooling power when scaling down.
y

∆T

x
L

z (a)

y′ = y / ϕ y

∆T

x′ = x / ϕ x
L′ = L / ϕ x

z′ = z / ϕ z
(b)
Figure 5.2.1: (a) An original standing-wave refrigerator in its original coordinate
system (b) a scaled-down standing-wave refrigerator in a scaled-down coordinate
system.
188 Chapter 5

The analysis is made through the following approach: starting point is from an
original refrigerator in original coordinates and a scaled-down refrigerator in
scaled-down coordinates. To make a general analysis, the three axes of the scaled-
down coordinate system are scaled with three different scaling factors: ϕ x , ϕ y , ϕ z ,
as shown in Fig.5.2.1.

Control
Q& H Control
volume
volume Q& H
TC TC
E& 2 E& 2
W&2in W&2in

TH TH
Q& C Q& C

(A) (B)

Figure 5.2.2: Two types of energy flow in a standing-wave: A and B.

TC TH

Driver

Stack Pressure
Pressure node Cold Hot anti-node
HX HX

Figure 5.2.2(a): Type A: a quarter-wave-length standing-wave cooler system.

Driver
TC TH

Pressure Stack Pressure


anti-node Cold anti-node
Hot
HX HX
Pressure node

Figure 5.2.2(b): Type B: a half-wave-length standing-wave cooler system.

The temperature span across the stack in the original refrigerator ∆T remains the
same in the scaled version. First, we assume that the stack and the two heat
Scaling considerations 189

exchangers are contained in a perfect resonator tube, which means that no


dissipation and no power flow is needed to maintain the standing wave going. So,
we can assume that all the dissipation of work flow takes place in the stack.
Generally speaking, there are two types of energy flow, as shown in Fig.5.2.2. We
name them as energy flow types A and B. In practical situations, type A and B are
illustrated in Figs. 5.2.2(a) and (b), respectively. The total energy flows are
differently composed in the two types. In type A, a control volume is applied to the
cold heat exchanger and part of the stack, as shown in Fig. 5.2.2 (A). The total
energy flow along the stack in type A is given by
E& 2 = Q& C . (5.2.1)
So, the cooling power in the cold heat exchanger is given by the total energy flow
from cold to hot heat exchanger.
Similarly, a control volume is applied in type B, containing the cold heat exchanger
and part of the stack. The energy balance in the control volumes is given by
E& 2 = W&2in + Q& C . (5.2.2)
W&2in is the work fed by the driver. In this case the cooling power equals the
difference between the total energy flow and the acoustic power input from the
driver.
It can be seen that the cooling power depends on the total energy flow along the
stack in both types. So, for type A, the cooling power is given by
Q& C = E& 2 , (5.2.3)
and for type B:
Q& C = E& 2 − W&2in , (5.2.4)
In order to investigate the variation of the cooling power after scaling down, it is
necessary to look at the variation of the total energy flow E& 2 along the stack.
Now the parameters of the scaled-down refrigerator system are applied to the
equations in chapter 2. The following equations are obtained:
resonator tube length L ′ = L
ϕx , (5.2.5)
angular frequency ω ′ = ϕ x ω , (5.2.6)
radian wave length D ′ = D
ϕx . (5.2.7)
In the stack region, now the following equations hold for the thermal penetration
depth of the fluid
2K ′ 2K 1
δ κ′ = = = δκ , (5.2.8)
ρ m′ c pω ′ ρ mc pϕ xω ϕx
the viscous penetration depth of the fluid
190 Chapter 5

2µ ′ 2µ 1
δν′ = = = δν , (5.2.9)
ρ m′ ω ′ ρ mϕ x ω ϕx
and the thermal penetration depth of the solid
2 K s′ 2Ks 1
δ s′ = = = δs . (5.2.10)
ρ s′csω ′ ρ s csϕ xω ϕx
While scaling, the ratio of stack spacing to thermal penetration depth is assumed to
be kept constant, because it is by definition an optimized parameter for a fixed ratio
y 0′ y
= 0 , (5.2.11a)
δ κ′ δ κ
it leads to
1
y 0′ = y0 , (5.2.11b)
ϕx
Also, the ratio between stack plate thickness and material thermal penetration depth
is assumed to be fixed while scaling:
l′ l
= , (5.2.12a)
δ s′ δ s
it leads to
1
l′ = l. (5.2.12b)
ϕx
In the original system, the perimeter is approximately calculated as follows:

Atube ∫∫S dydz


Π= = , (5.2.13)
y0 + l y0 + l
where Atube is the cross-sectional area of the tube.
In the scaled-down system, the perimeter is given by


Atube ∫∫ dy′dz′ (ϕ yϕ z ) −1 ∫∫ dydz
ϕx
Π′ = = S′ = S
= Π. (5.2.14)
y0′ + l ′ y0′ + l ′ ϕx −
1
2 ( y0 + l ) ϕ yϕ z
The Prandtl number is a material property of the working fluid, and is not affected
by scaling:
C pµ′ Cpµ
σ′ = = =σ . (5.2.15)
K′ K
The Rott’s functions f v′ , f κ′ and ε s′ for the scaled-down system are given by
substitution of equations (5.2.8) to (5.2.12ab) into equations (2.1.32) to (2.1.34).
These are also not affected by scaling:
Scaling considerations 191

y′
tanh (1 + i ) 0 ′  tanh (1 + i ) 0 
y
 δν   δν 
fν′ = = = fν , (5.2.16)
y ′ y
(1 + i ) 0 (1 + i ) 0
δν′ δν
y′
tanh (1 + i ) 0 ′  tanh (1 + i ) 0 
y
 δ κ   δ κ 
f κ′ = = = fκ , (5.2.17)
y ′ y
(1 + i ) 0 (1 + i ) 0
δ κ′ δκ

ρ m′ c pδ κ′ tanh (1 + i ) y0 δ ′  ρ mc pδ κ tanh (1 + i ) y0 δ 
ε s′ =  κ
=  κ
= εs . (5.2.18)
 
ρ s′csδ s′ tanh (1 + i ) δ ′ 
l ′ 
ρ s csδ s tanh (1 + i ) δ 
l 
 s  s

If the input acoustic pressure at the interface between the driver and the resonator
tube is constant while scaling, then, at the same positions relative to the resonator-
tube length, the acoustic pressure is also not affected (see appendix K)
p1′ = p1 . (5.2.19)
Therefore, the pressure and temperature gradient become
dp1′ dp1 dp
= = ϕx 1 ,
dx′ dx dx (5.2.20)
ϕx
dTm′ dTm dT
= = ϕx m .
dx ′ dx dx (5.2.21)
ϕx
Substitution of the equations (5.2.6) to (5.2.21) into the total energy flow equation
(2.1.72) yields

&

−1  ∏ y0
E2′ = (ϕ yϕ z ) ⋅ 
 dp
Im 
~  ~
p1 1 − fν −
(
fκ − fν
~


)
(1 + ε s )(1 + σ ) 
1

 2ωρ m  dx 

+
∏ y0 c p dTm dp1 dp ~
1
~
× Im  fν +
( )
fκ − fν (1 + ε s fν / fκ )  
~

2ω 3 ρ m (1 − σ ) dx dx dx  (1 + ε s )(1 + σ )  
− ϕ x (ϕ y ϕ z ) Π ( y 0 K + lK s ) m .
−1 dT
(5.2.22)
dx
The two terms scaling with different factors are defined as

E − acoustics =
∏ y0  dp
Im 
~  ~
p1 1 − fν −
fκ − fν
~
( 

)
(1 + ε s )(1 + σ ) 
1
2ωρ m  dx 

+
∏ y0 c p dTm dp1 dp~
1
~
× Im  fν +
( )
fκ − fν (1 + ε s fν / fκ ) 
~

(1 + ε s )(1 + σ ) 
,
2ω 3 ρ m (1 − σ ) dx dx dx
(5.2.23)

192 Chapter 5

and
dTm
E − conduction = ∏( y 0 K + lK s ) . (5.2.24)
dx
Eq. (5.2.22) can now be written as
ϕ
⋅ {E − acoustics} − x ⋅ {E − conduction}.
1
E& 2′ = (5.2.25)
ϕ yϕ z ϕ yϕ z
For energy flow type A, the cooling power equals the total energy flow along the
stack, as seen in Eq. (5.2.3)
ϕx
⋅ {E − acoustics} − ⋅ {E − conduction}.
1
Q& C′ = E& 2′ = (5.2.26)
ϕ yϕ z ϕ yϕ z
For energy flow type B, more work needs to be done. We assume that the input
acoustic pressure at the interface between the driver and the resonator tube p1in is
constant in scaling, i.e. p1′in = p1in , it is shown in appendix K that the volume
velocity at the interface U1in is obtained as:
p1′in
= (ϕ yϕ z ) ⋅ U1in .
p1in −1
U1′in = =
Z′0 (ϕ yϕ z )Z 0 (K.30)

The total input acoustic power at the interface of the driver and the resonator tube
into the resonator tube after scaling is given by:

2
[ ~
] −1 1

2
~
[ −1
]
W&2′in = ⋅ Re p1′in ⋅ U1′in = (ϕ yϕ z ) ⋅ ⋅ Re p1in ⋅ U1in = (ϕ yϕ z ) ⋅ W&2in
1
(5.2.27)

Therefore, for energy flow type B, according to Eq. (5.2.4), the cooling power in
the scaled-down system can be expressed as
ϕx
Q& C′ = E& 2′ − W&2′in =
1
ϕ yϕ z
{
⋅ ( E − acoustics) − W&2in − } ϕ yϕ z
⋅ {E − conduction}.

(5.2.28)
From Eq. (5.2.26) and (5.2.28), it is obvious that the cooling power in the scaled-
down system consists of two groups of energy flow scaling with different factors.
The first group of energy flow scales with a factor ϕ y ϕ z ( )
−1
, whereas the
conduction term scales as ϕ x ϕ y ϕ z ( )−1
. This means that the thermal conduction will
finally dominate the losses during scaling down. The cooling capacity of the
system will decrease after scaling down. So, there must be a limitation for scaling
down.
If the system scales down uniformly in all directions, i.e. ϕ x = ϕ y = ϕ z = ϕ , the
resultant conclusion is even more clear. Then Eq. (5.2.26) and (5.2.28) become
Q& C′ = ϕ −2 ⋅ {E − acoustics} − ϕ −1 ⋅ {E − conduction} (5.2.29)
and
Scaling considerations 193

{ }
Q& C′ = ϕ −2 ⋅ ( E − acoustics) − W&2in − ϕ −1 ⋅ {E − conduction}. (5.2.30)
So the cooling power decreases rapidly with ϕ −2 , but the conduction losses with
ϕ −1 showing that for large ϕ these will tend to dominate, making Q& C negative. So
for large ϕ and fixed ∆T the losses via conduction will be larger than the available
acoustic power.

Total energy flow density


The cross sectional area of the tube section where the stack is located
1
A= ∏(2 y 0 + 2l ) = ∏( y 0 + l ) . (5.2.31)
2
After scaling, the area is given by
ϕ x ( y0 + l )
= (ϕ yϕ z ) A .
−1
A′ = ∏ ′( y 0′ + l ′) = Π (5.2.32)
ϕ yϕ z ϕx
The total energy flow per unit of cross-sectional area after scaling is obtained as:

E& 2′  1
=
1  dp
Im 
~ 
1
p1 1 − fν −
(
~ Tm β fκ − fν 
~
) 
A′ 1 + l 2ωρ m  dx 
y0
(1 + ε s )(1 + σ ) 

+
1 cp dTm dp1 dp ~
1
~
× Im  fν +
( ) 
fκ − fν (1 + ε s fν / fκ )  
~

1 + l 2ω ρ m (1 − σ ) dx dx dx
y0
3
 (1 + ε s )(1 + σ )  

1  l  dT
−ϕx ⋅  K + K s  m . (5.2.33)
1+ l  y0  dx
y0
It can be rewritten as:
E& 2′  E − acoustics   E − conduction 
=  − ϕx ⋅  . (5.2.34)
A′  A   A 
The total energy flow per unit area decreases when the system scales down. The
energy flow density in the first term group remains the same, whereas the energy
flow per unit area due to gas and plate thermal conduction increases when scaling
down. Eq.(5.2.34) shows that the maximum scaling factor in x-direction is given by
the ratio of acoustic energy flow and conductive flow in the original case.

Coefficient of performance (COP)


With the assumption that the input acoustic pressure at the interface between the
driver and the resonator tube p1in is constant in scaling, i.e. p1′in = p1in , it is shown
in appendix K that the volume velocity at the interface U1in is obtained as:
194 Chapter 5

p1′in
= (ϕ yϕ z ) ⋅ U1in .
p1in −1
U1′in = =
Z′0 (ϕ yϕ z )Z 0 (K.30)

The total input acoustic power at the interface of the driver and the resonator tube
into the resonator tube after scaling is given in Eq.(5.2.27).
According to the definition, the COP of a standing-wave thermoacoustic
refrigerator after scaling is given as:
Q& C′
COP′ = . (5.2.35)
W&2′in
Therefore, substituting Eq. (5.2.26) and (5.2.27) into (5.2.35), the COP of the
scaled-down system of type A becomes

COP′ =
Q& C′
=
(ϕ yϕ z )−1 ⋅ {E − acoustics}− ϕ x (ϕ yϕ z )−1 ⋅ {E − conduction}
W&2′in (ϕ yϕ z )−1 ⋅W&2in
E − conduction
= COP − (ϕ x − 1) . (5.2.36a)
W& 2in
For type B, substitution of Eq. (5.2.28) and (5.2.27) into (5.2.35), the COP after
scaling becomes

COP′ =
(ϕ ϕ ) ⋅ {( E − acoustics ) − ∆W& }− ϕ (ϕ ϕ ) ⋅ {E − conduction}
y z
−1
2 x y z
−1

(ϕ ϕ ) ⋅W& y z
−1
2 in

E − conduction
= COP − (ϕ x − 1) . (5.2.36b)
W& 2in
From Eq. (5.2.36a) and (5.2.36b), it can be seen that the COPs for both types A and
B are described by exactly the same equation. The last term in Eq. (5.2.36ab) is the
product of the scaling factor and the original ratio of energy losses due to thermal
conduction of working gas and plates to acoustic work. This term causes the
reduction of COP after scaling. That means that the energy loss due to thermal
conduction of the working gas and plates will become more and more dominant,
when the system scales down.
For the cases of constant total energy flow E& 2 and constant total energy flow per
unit area E& 2 / A , explicit analytical expressions can not be obtained. These two
cases will be discussed later by numerical computation.
To make this theoretical analysis visible, a computation is made for uniform
scaling and for energy flow type A. Type B can be treated fully analogously. A
sketch of the reference system is given in Fig.5.2.3. It is assumed that the left end is
connected to a large buffer volume to work as an open end. The right end is driven
by a loudspeaker. Therefore, the acoustic power is fed into the resonator from the
end facing to the hot-end heat exchanger, as in type A of Fig. 5.2.2. The cold and
Scaling considerations 195

hot heat exchangers are assumed to be ideal. The configuration of the standing-
wave refrigerator described in Fig.3.3.2, Table 3.3.I (a) and (b) is used as the
reference system. It operates at 1000 Hz. The scaling performance of the original
configuration under different working conditions: mean pressure of 1 bar, 10 bar
and 100 bar is compared, while the drive ratio remains 10% all the time.
Comparison between different temperature span across the stack under the same
operational condition is also investigated. All these comparisons are made to study
how to expand the space for scaling down a specific system.

Lres
Pressure W& 2in
Pressure
node anti-node
Cold Hot
0 x HX HX
xstack=0.63Lres

Figure 5.2.3: A sketch of the reference standing-wave refrigerator.

Comparison a) Different temperature difference over the stack


(1) The stack position is fixed at x/Lres=0.63 (in the reference system that is x=0.16
m), the output cooling power Q& C is 13 W. The temperature difference over the
stack ∆Tstack is 100 K, mean pressure of 1 bar and drive ratio of 10%. Under the
condition that ∆Tstack is kept at 100 K, scaling down of the system leads to a change
of the energy group related to productive acoustic power Eq. (5.2.23) and the
thermal conduction of the working gas and stack plates Eq. (5.2.24). The energy
changes taking place at the leading end of the stack are shown in Fig. 5.2.4. The net
cooling power is the difference between the two curves in Fig.5.2.4 and that
decreases rapidly when the system size is reduced.
In Fig.5.2.4, the energy flow containing acoustic power, which is the first term
group in Eq. (5.2.26) and defined by Eq. (5.2.23), decreases faster than the
consumptive energy flow due to thermal conduction of the working gas and stack
plates. At some point, the two curves cross each other. If the scaling-down goes
further than that value, no cooling power is available anymore. Here in this specific
case, the value is approximately 45. That means that the model refrigerator can be
scaled down to around 1/45 of its original size, not smaller, if non-zero cooling
power is required. The decrease of cooling power at the same ∆Tstack after scaling-
down implies the reduction of cooling capability.
196 Chapter 5

1 E-acoustics
E-conduction

Log (energy flow) (W)


0

-1

-2

-3

-4

-5

1 10 100 1000
Scaling factor

Figure 5.2.4: Energy flows containing acoustic power and heat conduction through
working gas and stack plates as a function of the linear down-scaling factor.
Temperature difference over the stack remains constant (100 K) mean pressure is 1
bar, drive ratio=10%, x/Lres=0.63.

(2) If the system operates at another temperature span across the stack ∆Tstack, the
scaling performance will behave differently. Using the same configuration, and
placing the stack at the same position x/Lres=0.63 (x=0.16 m in reference system),
the output cooling power Q& C is increased to 39 W at a temperature difference over
the stack of 50 K. The mean pressure is 1 bar and drive ratio remains 10%. The
energy flows are plotted in Fig. 5.2.5.
In Fig.5.2.5, the two curves are similar to those in Fig.5.2.4. The energy flow
containing acoustic power decreases faster than the energy flow due to thermal
conduction of the working gas and stack plates. Again, at the crossing point of the
two curves the net cooling power decreases to zero. Here in this case, the
maximum scaling factor is around 177. That means that the reference refrigerator
can be scaled down to 1/177 of its original size, if we keep ∆Tstack at 50 K with
non-zero cooling power. Compared with the previous case, the factor to scaling-
down is greatly enlarged, however, under a lower performance condition. Further
scaling-down can be realized by the reduction of temperature difference over the
stack. This is of course not an attractive option because of the smaller and smaller
∆T. Another option is to increase the mean pressure of the system and keeping the
same drive ratio.
Scaling considerations 197

1 E-acoustics
E-conduction

Log (energy flow) (W)


0

-1

-2

-3

-4

-5
1 10 100 1000
Scaling factor

Figure 5.2.5: Energy flows containing acoustic power and heat conduction through
working gas and stack plates as a function of the linear down-scaling factor.
Temperature difference over the stack remains constant (50 K) mean pressure is
1bar, drive ratio=10%, x/Lres=0.63.

2
E-acoustics
E-conduction
1
Log (energy flow) (W)

-1

-2

-3

-4

1 10 100 1000
Scaling factor

Figure 5.2.6: Energy flows containing acoustic power and heat conduction through
working gas and stack plates as a function of the down-scaling factor. Temperature
difference over the stack remains constant (100K), mean pressure is 10 bar, drive
ratio=10%, x/Lres=0.63.
198 Chapter 5

Comparison b) Different mean pressure at the same drive ratio


Computations are made for the same system configuration working at mean
pressures 10 bar and 100 bar. The drive ratio is constant at 10%.
(3) If in the reference system, the mean pressure is 10 bar, and the stack position is
fixed at x/Lres=0.63 (x=0.16 m), the output cooling power Q& C is 121 W at a
temperature difference over the stack ∆Tstack of 100 K. The temperature over the
stack remains constant (100 K), when scaling down the system. The change of the
energy flows is shown in Fig.5.2.6. Based on the intersection point in Fig. 5.2.6,
the increase of the mean pressure leads to a much larger window for scaling-down.
In this case, it can be scaled down to nearly 1/387 of the original size, which is
about a factor of 10 more than in the 1 bar case of Fig.5.2.4.

(4) A much higher mean pressure 100 bar is applied to the same system
configuration for a further comparison. The drive ratio is 10%. The stack position
is fixed at x/Lres=0.63 (x=0.16 m in reference system) and the output cooling power
Q& C is 612 W in the reference system at the temperature difference over the stack
∆Tstack of 100 K. Keeping ∆Tstack constant (100 K), The change of the two groups
of energy flow by the scaling-down of the system is shown in Fig. 5.2.7.

2
E-acoustics
E-conduction
Log (energy flow) (W)

-1

-2

-3

-4

1 10 100 1000
Scaling factor

Figure 5.2.7: Energy flows containing acoustic power and heat conduction through
working gas and stack plates as a function of the linear down-scaling factor.
Temperature difference over the stack remains constant (100K), mean pressure
is100 bar, drive ratio=10%, x/Lres=0.63.
Scaling considerations 199

The intersection point in Fig 5.2.7 indicates that the system can be scaled to 1/1807
of the original size. Compared with case of mean pressure 10 bar, the window to
scale-down is increased again by increasing mean pressure.

5.2.2 Constant time-averaged total energy flow


As mentioned in section 5.2.1, for the analysis of ∆Tstack at constant energy flow
E& 2 and constant energy flow per unit area E& 2 / A , explicit analytical expressions
are not available. Therefore, only results via numerical computation are presented
here.
(1) The stack position is again fixed at x/Lres=0.63 (x=0.16 m in reference system),
and the output cooling power Q& C is 13 W (6593 W/m²) at the temperature
difference over the stack ∆Tstack of 100 K in the reference system. The mean
pressure is 1 bar and the drive ratio is 10%. At a constant total energy flow
E& 2 (13W), the scaling-down of the system leads to a decrease of the temperature
difference over the stack. It is shown in Fig 5.2.8. The curves for variation of
temperature difference over the stack ∆Tstack at other values of total energy flow are
also computed for comparison with the one of 13 W. The trend in Fig.5.2.8 is that
∆Tstack drops when the size of the system decreases. For each of the energy flows,
at some scaling factor, the temperature difference over the stack is zero. It indicates
the minimum size to scale down at a specific constant total energy flow along the
stack plates at which ∆T vanishes. For instance, the curve of E& 2 =1.3 W, the
system can be decreased by a factor of 7.2 maximum. If the total energy flow along
the stack becomes larger, the decrease in ∆Tstack becomes steeper. Looking at it
from another point of view, a smaller cooling power output enlarges the scaling-
down window for the same system.

(2) The same computation is carried out for the case of 10 bar mean pressure. The
stack position is fixed at x/Lres=0.63 (x=0.16 m in reference system), with an output
cooling power Q& C of 121 W at the temperature difference over the stack ∆Tstack of
100 K in the reference system. Keeping the energy flow E& 2 constant (121 W), the
scaling-down of the system leads to change of ∆Tstack. It is shown in Fig.5.2.9. As
seen from Fig.5.2.8 and 5.2.9, the performances of scaling-down for both cases are
similar.
200 Chapter 5

140

E2=13 W
120 E2=1.3 W
E2=0.13 W
100

80
∆Tstack (K)

60

40

20

0
0 3 6 9 12 15 18 21 24
Scaling factor

Figure 5.2.8: Temperature difference over the stack as a function of scaling factor.
Total energy flow remains constant. Mean pressure is1bar, drive ratio=10%,
x/Lres=0.63.

140
E2=121 W
120 E2=12.1 W
E2=1.21 W
100
∆Tstack (K)

80

60

40

20

0
0 3 6 9 12 15 18 21 24
Scaling factor

Figure 5.2.9: Temperature difference over the stack as a function of scaling factor.
Total energy flow remains constant, mean pressure is 10 bar, drive ratio=10%,
x/Lres=0.63.
Scaling considerations 201

5.2.3 Constant time-averaged total energy flow density

Using the same system configuration, the stack position is again fixed at
x/Lres=0.63 (x=0.16 m in reference system), and the output cooling power Q& C is 13
W at a temperature difference over the stack ∆Tstack of 100 K in the reference
system. The mean pressure of the system is 1 bar and drive ratio is 10%. Keeping
the energy flow density E& 2 / Ares constant (6593 W/m²), the scaling-down of the
system leads to a change of temperature difference over the stack. The behavior of
∆Tstack is shown in Fig.5.2.10. The curves for variation of the temperature
difference over the stack ∆Tstack at other values of energy flow density are also
plotted for comparison with the case of 6593 W/m².
For comparison with 1 bar, a mean pressure of 10 bar for the same configuration is
also considered and the resultant curves are plotted in Fig.5.2.11.
As shown in Fig. 5.2.10, ∆Tstack decreases as the system is scaled down, under the
prerequisite that the total energy flow density remains constant. By increasing the
constant energy flow density along the stack E& 2 / Ares , the maximum scaling factor
decreases. Furthermore, comparison of Figs.5.2.10 and 5.2.11 shows that
increasing the mean pressure leads to a better performance of the scaled-down
system.

2
100 E2/Ares=6593 W/m
2
E2/Ares=13186 W/m
2
E2/Ares=19780 W/m
80

60
∆Tstack (K)

40

20

0
1 10 100 1000
Scaling factor

Figure 5.2.10: Temperature difference over the stack as a function of scaling factor.
Total energy flow density remains constant. Mean pressure is 1 bar, and drive
ratio=10%, x/Lres=0.63.
202 Chapter 5

2
100 E2/Ares=59218 W/m
2
E2/Ares=118436 W/m
2
E2/Ares=177655 W/m
80

60
∆Tstack (K)

40

20

0
1 10 100 1000 10000
Scaling factor

Figure 5.2.11: Temperature difference over the stack as a function of scaling factor.
Total energy flow density remains constant. Mean pressure is 10 bar, and drive
ratio=10%, x/Lres=0.63.

It is important to note that in this case the scaling factor can be increased till
infinity without reaching a negative temperature difference. The reason for this is
obvious because on each curve the energy flow per surface area remains constant.
The decrease is purely due to thermal conduction caused by an increased thermal
gradient matching the enthalpy flow.

5.3 Traveling-wave systems

Like in the discussion of the standing-wave system, it is supposed that the


temperature difference across the regenerator remains unchanged while scaling
down. We focus on the variation of the cooling power. The reference traveling-
wave system has the configuration described in section 4.3, as shown in Fig.4.3.1.
Again, two coordinate systems are considered: an original refrigerator in an
original coordinate system and a scaled-down refrigerator in a scaled-down
coordinate system. As shown in Fig.5.2.1, the three axes of the scale-down
coordinate are scaled down with three different scaling factors: ϕ x , ϕ y , ϕ z .
Assuming that the diameters of the tube components are much larger than the
working-gas penetration depth, and that the mean pressure Pm remains unchanged,
Scaling considerations 203

the analysis for all the dimensions as made in section 5.2 from Eq. (5.2.5) to
(5.2.21) is the same for traveling-wave systems.
The traveling-wave system has the following parameters that are different from the
standing-wave system:
The low-Reynolds-number-limit flow resistance of the regenerator
R0′ ≈ 6 µ0′ Lreg
′ / Areg [
′ (rh′− reg ) 2 = 6 µ0 Lreg / ϕ x Areg (ϕ yϕ z ) −1 (rh − reg / ϕ x ) 2 ]
[ ]
= (ϕ yϕ z )6µ0 Lreg / Areg (rh − reg ) 2 = (ϕ yϕ z ) R0 , (5.3.1)
The porosity of the regenerator
′ = ψ reg ,
ψ reg (5.3.2)
The isothermal compliance
C0′ = ψ reg
′ Areg
′ Lreg
′ / pm′ = (ϕ xϕ yϕ z ) −1ψ reg Areg Lreg / pm = (ϕ xϕ yϕ z ) −1 C0 . (5.3.3)
Since the temperatures at the cold end and hot end remain the same, we have
τ ′ = TH′ / TC′ = τ . (5.3.4)
Therefore, we have for the earlier defined functions Eq.(F.21) and (F.22) in
appendix F:
b +2
(1 / τ ′) µ − 1
f ′(τ , bµ ) = = f (τ , bµ ) , (5.3.5)
(1 / τ ′ − 1)(bµ + 2)
 (1 / τ ′)bµ + 2 − 1 (1 / τ ′)bµ + 2 ln (1 / τ ′) 
g ′(τ , bµ ) =
1
 −  = g (τ , bµ ) . (5.3.6)
(1 / τ ′ − 1) 2  (bµ + 2)
2
bµ + 2 
The real wave number scales as follows:
k ′ = ω ′ / a′ = ϕ xω / a = ϕ x k . (5.3.7)
In combination with the length in the x direction Lx , we have
k ′Lx′ = ϕ x k ( Lx / ϕ x ) = kLx . (5.3.8)
Now substituting these in the complex coefficients from D1 to D4 (Eqs.(F.36a) to
(F.36d) in appendix F), we find that the scaled coefficients are:
[
D1′ = (1 − iω ′C0′ R0′ g ′(τ ′, bµ ) ) cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′fb2 sin(k ′L′fb ) sin(k ′Lcpl
′ ) / d cpl
′2 ]
πd ′fb2 R0′ f ′(τ ′, bµ )
+i
4 ρ m′ a′
[sin(k ′L′fb ) cos(k ′Lcpl′ ) + d cpl′2 cos(k ′L′fb ) sin(k ′Lcpl′ ) / d ′fb2 ]
= (1 − iωC0 R0 g (τ , bµ ) )[cos(kL fb ) cos(kLcpl ) − d 2fb sin(kL fb ) sin(kLcpl ) / d cpl 2
]
πd 2fb R0 f (τ , bµ )
+i
4ρm a
[sin(kL fb ) cos(kLcpl ) + d cpl
2
cos(kL fb ) sin( kLcpl ) / d 2fb ] = D1 ,

(5.3.9)
[
D′2 = − R0′ f ′(τ ′, bµ ) cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′ sin(k ′L′fb ) sin(k ′Lcpl
2
cpl
′ ) / d′ 2
fb ]
204 Chapter 5

4ρma
− i (1 − iω ′C0′ R0′ g ′(τ ′, bµ ) )
πd ′fb2
[sin(k ′L′fb ) cos(k ′Lcpl
′ )

+ d ′fb2 cos(k ′L′fb ) sin( k ′Lcpl


′ ) / d cpl
′2 ] = (ϕ ϕ )D
y z 2, (5.3.10)
iω ′C0′ ln τ ′
D′3 =
1−τ ′
[
cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′fb2 sin(k ′L′fb ) sin(k ′Lcpl
′ ) / d cpl
′2 ]
iπd ′fb
[ ]
2

− sin(k ′L′fb ) cos(k ′Lcpl


′ ) + d cpl
′2 cos(k ′L′fb ) sin(k ′Lcpl′ ) / d ′fb2
4 ρ m aτ ′
= (ϕ yϕ z ) D3
−1
(5.3.11)
[
D′4 = cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′ sin(k ′L′fb ) sin(k ′Lcpl
2
cpl
′ ) / d ′ /τ ′ 2
fb ]
4 ρ aω ′C ′ ln τ ′
+ m 2 0
πd ′fb (1 − τ ′)
[
sin(k ′L′fb ) cos(k ′Lcpl
′ ) + d ′fb2 cos(k ′L′fb ) sin(k ′Lcpl
′ ) / d cpl
′2 ]
= D4 . (5.3.12)
Substitution of these scaled coefficients into Eq. (4.3.33) yields:
4 ρ m aD′4
D′2 cos(k ′Ltb′ ) − i sin(k ′Ltb′ )
πdtb′2
Z′fb = = (ϕ yϕ z )Z fb . (5.3.13)
4 ρ aD′
1 - D1′ cos(k ′Ltb′ ) + i m 2 3 sin(k ′Ltb′ )
πdtb′
Substitution of the coefficients into Eq. (4.3.36) yields:
 4 ρ aD′ 
Z′input =  D′2 cos(k ′Ltb′ ) − i m 2 4 sin(k ′Ltb′ ) [1 + D1′D′4 − D′2 D′3 −
 πd tb′ 
−1
 πd tb′2 D′2 4 ρ m aD′3  
′ ′ ′ ′
(D1 + D4 ) cos(k Ltb ) + i + 
 sin( k ′L′ )  = (ϕ yϕ z )Z input .
 4 ρ m a π d ′
tb
2

tb

(5.3.14)
Via substitution of Eq. (5.3.14) into (4.3.37), the scaled-down acoustic impedance
of the refrigerator system becomes:
4ρma
Z′input cos(k ′Lres
′ )+i sin(k ′Lres
′ )
4ρma ′2
πd res
Z′a − rfg = = (ϕ yϕ z )Z a − rfg . (5.3.15)
πd res
2
4ρma
cos(k ′Lres
′ ) + iZ′input sin(k ′Lres
′ )
′2
πd res
Substitution of these scaled-down parameters into the equations for real
coefficients θ1 to θ 4 (Eqs.(F.37a) to (F.37d) in appendix F) leads to:
θ1′ = cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′fb2 sin(k ′L′fb ) sin( k ′Lcpl
′ ) / d cpl
′2 = θ1 , (5.3.16a)
θ 2′ = cos(k ′L′fb ) cos(k ′Lcpl
′ ) − d ′ sin(k ′L′fb ) sin(k ′Lcpl
2
cpl
′ ) / d ′ = θ2 ,
2
fb (5.3.16b)
θ3′ = sin( k ′L′fb ) cos(k ′Lcpl
′ ) + d ′ cos(k ′L′fb ) sin(k ′Lcpl
2
cpl
′ ) / d ′ = θ3 ,
2
fb (5.3.16c)
Scaling considerations 205

θ 4′ = sin(k ′L′fb ) cos(k ′Lcpl


′ ) + d ′fb2 cos(k ′L′fb ) sin(k ′Lcpl
′ ) / d cpl
′2 = θ 4 . (5.3.16d)
Substitution of the above scaled-down parameters into the Eqs.(4.3.53), (4.3.54)
and (4.3.45), for coefficients of Θ1 , Θ 2 and Θ3 yields
2
 4 ρ a sin(k ′L′ )  4 ρ a sin(2k ′L′ )  1 
Θ1′ = (cos(k ′Lres
′ )) +  m 2 res 
+ m  = Θ1 ,
2 res
Im 
 πd res
′ Z′a − rfg  π ′2
d res 
 Z′a − rfg 
 
(5.3.17)

Θ′2 =
ψ reg
′ Areg
′ R0′  θ ′θ ′ θ ′θ ′ 
Im  ~ + 1 2 3 4 1 − fν′ −
(
~ Tm β fκ′ − fν′  )~

2ω ′ρ m Lreg ′  Z′fb

Z′fb 

(1 + ε s′ )(1 + σ ′) 

+
′ Areg
ψ reg ′ R0′  πd ′fb2θ1′θ3′ 4 ρ m aθ 2′θ 4′   ~ Tm β fκ′ − ~
+  − ′ −
( )fν′ 
(1 + ε s′ )(1 + σ ′)
Re 1 f
2ω ′ρ m Lreg ′  4 ρ m a π d
2 
′fb2 Z′fb  
ν
 
′ c pT0 R0′2 1 / τ ′ − 1  θ 2′2  πd ′fb2θ 3′  πd ′fb2θ 2′θ 3′ Im[Z′fb ] 
2
ψ reg
′ Areg
− +  +
2ω ′3 ρ m (1 − σ ′) Lreg
′2 Lreg ′  Z′ 2  4 ρ m a  2ρma 2 
Z′fb 
 fb
~

⋅ Im  fν +
( )
fκ′ − fν′ (1 + ε s′ fν′ / fκ′ ) 
~
 = (ϕ yϕ z ) Θ 2 ,
−1

 (1 + ε s′ )(1 + σ ′)  (5.3.18)
~ ~ ~
Re[D′2 D′4 ] D′2 D′3 D1′ D′4
+ ~ ] = (ϕ yϕ z ) Θ3 .
~ −1
Θ′3 = Re[D1′ D′3 ] + + Re[
Z′fb
2
Z ′fb ′
Z fb (5.3.19)

After substituting these scaled coefficients into Eq. (4.3.52), the scaled-down
cooling power is given by:
1 − (1 / τ ′)
Q& C′ = p1′in Θ1′ {Θ′3/ 2 + Θ′2 } − ψ reg
2
[
′ K + (1 − ψ reg
′ ) K s Areg
′ T0 ] ′
Lreg
p1′in
2
ϕx
=
1
p1in Θ1 ⋅ {Θ3 / 2 + Θ 2 }−
2
[ψ K + (1 −ψ reg ) K s ]
p1in ϕ yϕ z ϕ yϕ z reg
2

1 − (1 / τ )
⋅ AregT0 (5.3.20)
Lreg
If the input acoustic pressure from the driver remains the same, i.e. p1′in = p1in , then
Eq. (5.3.20) becomes
ϕx 1 − (1 / τ )
Q& C′ =
ϕ yϕ z
1
p1in Θ1{Θ3 / 2 + Θ 2 } −
2

ϕ yϕ z
[
ψ reg K + (1 −ψ reg ) K s ]AregT0
Lreg
.

(5.3.21)
206 Chapter 5

As seen from Eq. (5.3.21), the cooling power after scaling down consists of two
groups of terms scaling with different factors. Similar to the standing-wave system
in section 5.2 for, the two groups of terms are defined as
E − acoustics = p1in Θ1 ⋅ {Θ3 / 2 + Θ 2 },
2
(5.3.22)
1 − (1 / τ )
[
E − conduction = ψ reg K + (1 − ψ reg ) K s AregT0 ] Lreg
. (5.3.23)

Eq. (5.3.21) can be rewritten as:


ϕ
⋅ {E − acoustics} − x ⋅ {E − conduction} .
1
Q& C′ = (5.3.24)
ϕ yϕ z ϕ yϕ z
The cooling power of the original system is given by:
Q& C = {E − acoustics} − {E − conduction} (5.3.25)
From Eq. (5.3.24), it is obvious that the cooling power in the scaled-down system
consists of two groups of energy flow scaling with different factors. The first group
of energy flow scales with a factor ϕ y ϕ z ( )
−1
, while the conduction term scales
(
as ϕ x ϕ y ϕ z )
−1
. This is fully analogous to the standing-wave system Eq.(5.2.26).
The scaling behaviour is also fully analogous to that of the standing-wave system.

Coefficient of performance (COP)


For the computation of the COP the expression is needed for Θ′4 (Eq.(4.3.58)) that
is in the scaled version given by:
 ~
 4 ρ m aD′3  ~
Θ′4 = Re (1 - D′4 cos(k ′Ltb′ ) )1 − i
 sin(k ′Ltb′ )  + D1′ (D′4 - cos(k ′Ltb′ ) )
  πdtb′2

~
πd ′2 D′  ~ 4 ρ aD′ 
− i tb 2 sin(k ′Ltb′ ) /  D′2 cos(k ′Ltb′ ) + i m 2 4 sin(k ′Ltb′ )  = (ϕ yϕ z ) Θ 4
−1

4ρma   πdtb′ 


(5.3.26)
Assuming p1′in = p1in , the scaled-down COP is obtained by substitution of all the
scaled-down parameters into Eq. (4.3.60)

COP′ =
Θ′3 + 2Θ′2 2 ψ reg

[
′ K + (1.0 − ψ reg
′ ) K s Areg ]
′ T0 1 − (1 / τ ′)

Θ′4 p1′in Θ1′Θ′4
2

Lreg

=
Θ3 + 2Θ 2
− ϕx
[ ⋅
]
2 ψ reg K + (1.0 − ψ reg ) K s AregT0 1 − (1 / τ )
Θ4 p1in Θ1Θ 4
2
Lreg

= COP − (ϕ x − 1)
[ ]
2 ψ reg K + (1.0 − ψ reg ) K s AregT0 1 − (1 / τ )
⋅ . (5.3.27)
p1in Θ1Θ 4
2
Lreg
Scaling considerations 207

By substituting Eq. (5.3.24), (4.3.59), (5.3.26), (5.3.17) into definition that


COP′ = Q& C′ / W&in′ , the COP of the scaled-down system becomes:
Q& C′ (ϕ yϕ z ) ⋅ {E − acoustics} − ϕ x (ϕ yϕ z ) ⋅ {E − conduction}
−1 −1

COP′ = =
W&in′ (ϕ ϕ )−1 ⋅W&
y z in

E − conduction
= COP − (ϕ x − 1) . (5.3.28)
W&in

Eq. (5.3.27) or (5.3.28) shows that the COP for the scaled system decreases. The
last term in Eq. (5.3.27) or (5.3.28) is the product of the scaling factor and the
original ratio of energy losses due to thermal conduction of working gas and
regenerator material to acoustic work. This term causes the reduction of COP after
scaling. This is also fully analogous to the standing-wave system Eq.(5.2.36).

As an example, scaling is considered of the traveling-wave refrigerator designed


and used in section 4.5. The configuration is described in section 4.5.3 and shown
in Fig.4.5.1, i.e. the original system. For simplicity, the computation is made for a
uniform scaling, i.e. a universal scaling factor for all three dimensions:
ϕ x = ϕ y = ϕ z = ϕ . The scaling performance of the original configuration under
different working conditions: different temperature spans across the regenerator at
the same mean pressure of 11 bar are compared. Comparison between different
mean pressures, 11 bar and 15 bar, with constant temperature span across the
regenerator (40 K) under the same operational condition (the drive ratio remains
2.5% all the time) is also investigated. These comparisons are made to study how
to expand the space for scaling down a specific system.

Comparison a) Different temperature difference over the regenerator


(1) The reference-case traveling-wave refrigerator works at 320 Hz and the output
cooling power Q& C is 33 W. The temperature difference over the regenerator is 40 K,
and the mean pressure is 11 bar. Under the condition that the temperature
difference over the regenerator remains constant at 40 K, scaling of the system
leads to a change of the energy flow defined by Eq. (5.3.22) and the heat
conduction of the working gas and regenerator material Eq. (5.3.23). This is shown
in Fig. 5.3.1. The net cooling power, i.e. difference between the curves in Fig.5.3.1,
decreases fast when the system decreases in size.
As seen in Fig.5.3.1, the scaling behaviour is fully analogous to that of standing-
wave system in section 5.2: the energy flow defined by Eq. (5.3.22), decreases
faster than the consumptive energy flow due to thermal conduction of the working
gas and regenerator material. At a scaling factor of about 4, the two curves cross.
208 Chapter 5

That means that the model refrigerator can be scaled down to around one fourth of
its original size, if non-zero cooling power is required.

2.0

1.5 E-acoustic
E-conduction
Log(energy flow) (W)

1.0 Pmean=11 bar, ∆T=40 K

0.5

0.0

-0.5

-1.0

-1.5

1 10 100
Scaling factor

Figure 5.3.1: Energy flows containing acoustic power and heat conduction through
working gas and regenerator material as a function of the linear down-scaling
factor. Temperature difference over the stack remains constant (40 K) mean
pressure is 11 bar.

(2) If the system operates at a smaller temperature span across the regenerator, the
scaling performance will behave similarly but the scale factor can be larger. When
the temperature difference across the regenerator is 23 K, the output cooling power
Q& C is increased to 37 W. It still works at 320 Hz and the mean pressure remains 11
bar. The energy flows with scaling are plotted in Fig. 5.3.2.
In Fig.5.3.2, the two curves decrease differently and cross at around 9. That
indicates that the reference refrigerator can be scaled down to 1/9 of its original
size, if we keep the temperature difference over the stack at 23 K with non-zero
cooling power. Compared with the last case, the space for scaling-down is enlarged
under a lower temperature difference over the regenerator. Analogous to the
discussion of standing-wave systems in section 5.2, further scaling-down can be
realized only at the cost of a reduced temperature difference over the regenerator.
Scaling considerations 209

2.0

E-acoustic
1.5
E-conduction
Pmean=11 bar, ∆T=23 K

Log(energy flow) (W)


1.0

0.5

0.0

-0.5

-1.0

-1.5
1 10 100
Scaling factor

Figure 5.3.2: Energy flows containing acoustic power and heat conduction through
working gas and regenerator material as a function of the linear down-scaling
factor. Temperature difference over the stack remains constant (23 K) mean
pressure is 11 bar.

2.0
E-acoustic
1.5 E-conduction
Pmean=15 bar, ∆T=40 K
1.0
Log(energy flow) (W)

0.5

0.0

-0.5

-1.0

-1.5
1 10 100
Scaling factor

Figure 5.3.3: Energy flows containing acoustic power and heat conduction through
working gas and regenerator material as a function of the linear down-scaling
factor. Temperature difference over the stack remains constant (40 K) mean
pressure is 15 bar.
210 Chapter 5

Comparison b) Different mean pressure at the same drive ratio


Computations are made for the system working at mean pressures of 11 bar and 15
bar. The drive ratio is constant at 2.5%.
(3) In reference system, at a mean pressure of 15 bar, the output cooling power
Q& C is 45 W at the temperature difference over the regenerator of 40 K. The change
of the energy flows under scaling at constant temperature difference is shown in
Fig.5.3.3. Based on the intersection point in Fig. 5.3.3, the increase of mean
pressure (drive ratio constant) leads to a slightly bigger window for scaling-down
in comparison with case (1) in Fig.5.3.1. In this case, it can be scaled down to
nearly 1/5 of the original size.

5.4 Conclusions

In the case of constant temperature difference over the stack or the regenerator, the
comparison between Eq. (5.2.26) or (5.2.28) for standing-wave systems and Eq.
(5.3.24) for traveling-wave shows that the scaling behaviour of a standing-wave
system is the same as that of a traveling-wave system. The cooling power in the
scaled-down system consists of two groups of energy flow scaling with different
factors. The first group of energy flow is related to the acoustic power and scales
with a factor (ϕ yϕ z ) −1 , whereas the conduction term scales as ϕ x (ϕ yϕ z ) −1 . This
means that the thermal conduction will finally dominate the losses during scaling
down. The cooling capability of the system will decrease when scaling down.
Therefore, there is a limitation for scaling down, when the temperature difference
of the scaled-down system becomes zero.
The COP decreases rapidly under scaling, as can be seen from Eq. (5.2.36) for a
standing-wave system and Eq. (5.3.28) for a traveling-wave system. The term,
which is the product of the scaling factor and the original ratio of energy losses due
to thermal conduction to acoustic work, causes the reduction of COP after scaling.
That means that the energy loss due to thermal conduction will become more and
more dominant, when the system scales down.
Chapter 6

Conclusions and recommendations

6.1 Conclusions

Chapter 2
In chapter 2 the general analytical expressions are derived for thermoacoustic
systems. Here the main definitions of the functions are made that are used in
chapters 3 and 4. The function E& 2 that describes the time-averaged total energy
flow in a stack plays an important role in the scalebility of the thermoacoustic
systems.

Chapter 3:
In chapter 3 a standing wave system is modeled using analytical expressions for
cooling power and COP. This model describes clearly the temperature distribution
in a stack of a tubular standing wave device driven by a loudspeaker. The influence
of the stack plate spacing is studied in detail, as well as the position dependency of
the stack. It is clear that the stack plate spacing must be sufficiently small to have
highest performance of heat transport. The model is compared with experimental
data of a standing-wave “thermoacoustic couple” that could be positioned in the
resonator. The model predicts correctly the trends, although additional losses (that
are not modeled) lead to discrepancy between the model results and measurements.
It is clear that a correct choice of the total energy flow E& 2 and loss function plays
an important role in matching experiment and model.

Chapter 4
With the analytic model that is derived in chapter 4 traveling wave systems can be
modeled. The great advantage of this model is that it is a fast solver because a full
analytic expression is obtained to model the complete traveling wave acoustic
feedback phenomenon of the apparatus. Therefore it is easy to change sizes and
212 Chapter 6

lengths of the geometry in the model and run the model to obtain performance
curves.
There is a good agreement between the model derived and developed in this thesis,
and the Los Alamos DeltaE solver, when the simulation results of both models
applied to Swift's traveling-wave engine are compared. Therefore the model
developed in this thesis was used to simulate, design and optimize a new device
built at TU/e.
The model has made clear that optimization of an around 1.3 meter size traveling-
wave cooler, driven by an external driver (loudspeaker) is possible without using a
compliance. The optimum system design was attained for a one-diameter-size
feedback tube containing heat exchangers and regenerator.
Measurements on different regenerator materials in a thermally driven coaxial
traveling-wave apparatus have made clear that the MESH number (for wire gauzes)
and CPSI number (for honeycombs) are very sensitive parameters. The maximum
performance of the system that was tested in this thesis occurred at a (hydraulic
radius)/(thermal penetration depth) ratio of 0.30 for stainless steel wire gauzes and
0.16 for honeycombs (the reference thermal penetration depth taken at average
regenerator temperature of 490 K). Besides that the performance depends also on
the porosity of the regenerator. Maximum performances will generally occur at
regenerator porosities between 80-90%. These sensitivities are in agreement with
predictions by the model developed in this thesis. The different regenerators with
different porosity and hydraulic radius in the measurements result in different
acoustic impedances globally and locally, which finally result in different phasing
between the pressure and the velocity in the regenerator and thus lead to different
thermodynamic cycles. That makes a difference in the final performance of the
engine. The experiments have shown that wire gauze materials exceed the
performance of honeycombs with a factor 2. Probably this is due to two effects.
One effect is the difference in thermal capacity between the light ceramic
honeycombs and the metal wire gauze. A second reason might be the advantage of
the randomized structure of wire gauze regenerators. The randomized nature of
these wire gauzes leads to a more efficient thermal contact with the oscillating gas
even though at higher porosity of the wire gauze regenerator itself in comparison
with the honeycombs.
The experiments with the traveling wave cooler have demonstrated that a 1.3-
meter-long (small) traveling wave system cools without compliance. So
thermoacoustic systems can be designed without the special precaution of
compliance. Further scaling down seems possible.
The thermal performance that was measured in the traveling wave cooler is about a
factor 2 lower in comparison with the model results. The trends of the model
Conclusions and recommendations 213

compare very well with the measurements. The difference between experiment and
design model is due to the neglect of most of losses, which take place inside the
system in practical operations, in the analytical model. This is probably caused by
streaming, which can only be removed by inserting additional flow straightners.
The experiments have demonstrated that it is indeed possible to manufacture
efficient heat exchangers from copper fins, using high end spark cutting technology.
This manufacturing is a tedious task of high tech machining, and is very expensive.
Downsizing to smaller coolers will obviously push up the limits of making even
smaller heat exchange components and stacks for which probably lithography
technology is needed.

Chapter 5
The findings of chapter 5 show that the cooling capacity, and the COP decrease
rapidly when scaling down. Therefore, there is a limitation for scaling down.
It is possible to make a mini-thermoacoustic standing wave machine, although
there is always a thermal conduction balance limitation for scaling down. For a
standing-wave system, the effective cooling power decreases rapidly when scaling
down with the prerequisite that the temperature span over the stack remains
constant (see Fig. 5.2.4). Although the reference system can be only scaled down to
nearly 1/45th of the original size before this thermal conduction loss balance occurs,
the further scaling down can be realized by taking some methods. The methods are:
(1) reducing the required constant temperature span over the stack (see Fig. 5.2.5);
(2) increasing filling pressures and applying the same relative pressure ratio (see
Figs. 5.2.6 and 5.2.7). If the filling pressure is as high as 100 bar, the scaling down
system can be as small as 1/1800th of the original size (see Fig.5.2.7), which is
around 0.142 mm long system and small enough for many applications. If the
prerequisite is keeping constant time-averaged total energy flow or constant time-
averaged total energy flow density, the temperature span over the stack ∆T
decreases when scaling down (see Figs. 5.2.8, 5.2.9, 5.2.10 and 5.2.11).
For traveling wave systems approximately the same rules apply as for standing
wave systems. The reference system in chapter 5 having a length in the order of
one meter was taken as our own model system which is already a relatively small
traveling wave system. The scaling analysis shown with figures 5.3.1 and further
makes clear that a factor 10 for a 11 bar system is the maximum scale down factor.
This means that it is impossible to scale these systems down below a size of 10 cm.
Again using very high average pressures of 100 bar or more, but in that case
extremely high power drivers will be needed. This seems not compatible or at least
a big challenge.
214 Chapter 6

The findings of chapter 5 tell us in general that the COP will definitely deteriorate
at smaller devices. This limitation of scaling down is caused by the conduction
losses. Therefore, the exploration on effective thermoacoustic materials for stack,
and regenerator with less conduction losses is recommended by the author.

6.2 Recommendations

In order to validate better, and improve the standing wave model, a standing wave
system is needed with warm and cold heat exchanger. This assures a fixed thermal
boundary condition at the end of the stack, whereas in the tested system in this
thesis (thermoacoustic couple) the boundary conditions are floating.
The scalability should be tested by building a small scale standing wave device
with a length of about 10 cm, this will prove experimentally the statements in this
thesis.
The performance of honeycombs should be explored more than was done in this
thesis with the coaxial traveling wave system. Emphasis should be placed on thin
walled glass stack materials with porosities of 90%. Here even care could be taken
of cell size versus temperature, meaning a tapered regenerator where penetration
depth is always matched with the correct cell size at variable temperature along the
regenerator. It is possible to manufacture ultra thin walled honeycombs from Pyrex
glass.
The mechanically driven traveling wave cooler that was tested in chapter 4 can be
improved by placing additional thermometers in the regenerator in order to
investigate the temperature gradient, and to find out if streaming takes place in the
regenerator itself. Furthermore the effect of flow straightners should be explored in
order to diminish the heat leak between the cold side of the regenerator and the
room temperature heat exchanger.
The performance of the multi-microphone system should be investigated, and
improved. This can partially be done by isolating the cooler with a flexible bellows
from the driver, or by isolating the driver itself so that the effect of mechanical
vibrations can be diminished. With a well working multi-microphone measurement
system the acoustical power can be compared with the power measured from the
heat exchangers. Another way to improve the multi-microphone measurement is to
increase the distance between the microphones so that a larger phase difference
will occur resulting in more accurate measurements.
The Fortran design model should be used to make optimization studies of traveling
wave systems. When the current experimental mechanically driven cooler is
improved the new results should be compared again with the model. It is expected
that model and experiment will then be in better agreement with each other.
Appendices

A: momentum equation derivation


In chapter 2, the momentum equation was derived as
dp 1 ∂ 2 u1
iωρ m u1 = − +µ . (2.1.19)
dx ∂y 2
Rewriting Eq. (2.1.19) into a standard form as an ordinary second order differential
equation gives
∂ 2 u 1 iωρ m 1 dp 1
− u1 = . (A.1)
∂y 2
µ µ dx
The solution u1 of this equation can be written as the sum of the solution of the
homogenous equation and a particular solution. The homogenous equation is
∂ 2 u 1 iωρ m
− u1 = 0 . (A.2)
∂y 2 µ
Assuming a solution of the form
u1 = Ce αy , (A.3)
and substitution of Eq.(A.3) into (A.2) yields
iωρ m
α2 − = 0. (A.4)
µ
Thus, the two roots of Eq. (A.4) are
1+ i
α1, 2 = ± . (A.5)
δv
The viscous penetration depth is defined in Eq. (2.1.14) as δ v = 2 µ ρ mω .
Therefore, the solution of the homogenous equation (A.2) is obtained
u1 = C1 ⋅ e (1+i ) y / δ v + C 2 ⋅ e − (1+i ) y / δ v . (A.6)
To obtain the particular solution, it can be assumed that
u1* = C1 ( y ) ⋅ e(1+ i ) y / δ v + C2 ( y ) ⋅ e − (1+ i ) y / δ v . (A.7)
The unknown coefficient functions C1 ( y ) and C2 ( y ) for the particular solution
can be found by solving the following joint equations
216 Appendix

dC1 ( y ) (1+ i ) y / δ v dC2 ( y ) − (1+ i ) y / δ v


⋅e + ⋅e =0
dy dy
(A.8)
dC1 ( y ) (1+ i ) y / δ v 1 + i dC2 ( y ) − (1+ i ) y / δ v  1 + i  1 dp1
⋅e ⋅ + ⋅e ⋅  −  = ⋅
dy δv dy  δ v  µ dx
Solving the two joint equations, the solutions are
dC1 ( y ) 1 dp1 δ v − (1+ i ) y / δ v
= ⋅ ⋅ e , (A.9)
dy 2 µ dx 1 + i
dC2 ( y ) − 1 dp1 δ v (1+ i ) y / δ v
= ⋅ ⋅ e . (A.10)
dy 2 µ dx 1 + i
Integration of Eq. (A.9) and (A.10) yields
idp1 − (1+ i ) y / δ v
C1 ( y ) = ⋅e⋅ , (A.11)
2 ρ mω dx
i dp
C2 ( y ) = ⋅ 1 ⋅ e (1+ i ) y / δ v . (A.12)
2 ρ mω dx
Substitution of Eq. (A.11) and (A.12) into Eq. (A.7) yields the particular solution
i dp1
u1* = ⋅. (A.13)
ρ mω dx
The summation of solutions of the homogenous equation and the particular
solution yields the final general solution:
i dp 1
u1 = C1 ⋅ e (1+i ) y / δ v + C 2 ⋅ e −(1+i ) y / δ v + ⋅ . (A.14)
ρ mω dx
Next, the boundary conditions are applied to obtain the coefficients C1 and C 2 .
Boundary conditions:
∂u 1
a) at y = 0 , because of the symmetry, =0
∂y
b) at y = y 0 , because of the solid wall, u1 = 0 .
By using boundary condition a, we obtain
C1 = C 2 (A.15)
By using boundary condition b, the coefficients are obtained
−i dp1 1
C1 = ⋅⋅ . (A.16)
ρ mω dx 2 cosh[(1 + i ) y 0 / δ v ]
This is the description of the oscillatory velocity profile as dependant on the
oscillatory pressure gradient including viscous terms.
i dp1  cosh[(1 + i ) y δ v ] 
u1 = 1 − .
ωρ m dx  cosh[(1 + i ) y 0 δ v ] 
(A.17)
Appendix 217

B: derivation of the temperature of the solid


plate
In chapter 2, the equation of the temperature of the solid plate was derived as
∂ 2 Ts1
iωTs1 = κ s . (2.1.23)
∂y ′ 2
Rewrite it as
∂ 2 Ts1 iω
− Ts1 = 0 . (B.1)
∂y ′ 2 κ s
It is a homogenous equation. Assume a solution of the form
Ts1 = Ce αy′ . (B.2)
Substitution of Eq. (B.2) into (B.1) yields

α2 − = 0. (B.3)
κs
The two roots of Eq. (B.3) are
1+ i
α1, 2 = ± . (B.4)
δs
Therefore, the solution of the homogenous equation (B.1) is obtained
Ts1 = C1 ⋅ e (1+i ) y ′ / δ s + C 2 ⋅ e − (1+ i ) y′ / δ s . (B.5)
Next, the boundary conditions are applied to get the coefficients C1 and C 2 .
Boundary conditions:
∂Ts1
c) at y ′ = 0 , because of the symmetry, = 0;
∂y ′
d) at y ′ = l , because of the solid wall, Ts1 = Tb1 , where Tb1 is temperature
amplitude at the boundary, and yet undetermined with the temperature of
the fluid together.
By using boundary condition c, we obtain
C1 = C 2 (B.6)
By using boundary condition d, the coefficients are obtained
Tb1
C1 = . (B.7)
2 cosh[(1 + i )l / δ s ]
Substitution of Eq. (B.6) and (B.7) into Eq. (B.5) yields the final solution of Eq.
(B.1)
cosh[(1 + i ) y ′ δ s ]
Ts1 = Tb1 .
cosh[(1 + i )l δ s ]
(B.8)
218 Appendix

C: derivation of the temperature oscillation of


the fluid layer
In chapter 2, the heat transfer equation was reduced to
 dTm  ∂ 2T
ρ m c p  iωT1 + u1  − iωp1 = K 21 . (2.1.29)
 dx  ∂y
Rewrite Eq. (2.1.29) into a standard form as an ordinary second order differential
equation
∂ 2T1 iωρ mc p ρ c dT iω
− T1 = m p m u1 − p1 . (C.1)
∂y 2
K K dx K
As did in the above part, the solution T1 of this equation can be written as the sum
of the solution of the homogenous equation and a particular solution. First, we start
with solving the homogenous equation of Eq. (C.1)
∂ 2 T1 iωρ m c p
− T1 = 0 . (C.2)
∂y 2 K
Assume a solution of the form as
T1 = Ce αy . (C.3)
Substitution of Eq. (C.3) into (C.2) yields
iωρ m c p
α2 − = 0. (C.4)
K
The two roots of Eq. (C.4) are
1+ i
α1, 2 = ± , (C.5)
δκ
where the fluid’s thermal penetration depth is
2K
δκ = . (C.6)
ρ m c pω
Therefore, the solution of the homogenous equation (C.2) is obtained
T1 = C1 ⋅ e (1+i ) y / δ κ + C 2 ⋅ e −(1+i ) y / δ κ . (C.7)
To obtain the particular solution, assume the particular solution is of the expression
as
T1* = C1 ( y ) ⋅ e(1+ i ) y / δ κ + C2 ( y ) ⋅ e − (1+ i ) y / δ κ . (C.8)
The unknown coefficient functions C1 ( y ) and C2 ( y ) for the particular solution
can be found by solving the following joint equations
dC1 ( y ) (1+ i ) y / δ κ dC2 ( y ) − (1+ i ) y / δ κ
⋅e + ⋅e =0
dy dy
Appendix 219

dC1 ( y ) (1+ i ) y / δ κ 1 + i dC2 ( y ) − (1+ i ) y / δ κ  1+ i  iωp1


⋅e ⋅ + ⋅e ⋅  −  = −
dy δκ dy  δκ  K
ic p dTm dp1  cosh[(1 + i ) y / δ v ] 
+ 1 − 
ωK dx dx  cosh[(1 + i ) y0 / δ v ] 
(C.9)

Solving the two joint equations, the solutions are


dC1 ( y ) 1 δ κ − (1+ i ) y / δ κ  ic p dTm dp1  cosh[(1 + i ) y / δ v ]  iω 
= e  1 −  − p1 
dy 2 1+ i  ω K dx dx  cosh[(1 + i ) y 0 / δ v ]  K 
(C.10)
dC2 ( y ) − 1 δ κ (1+ i ) y / δ κ  ic p dTm dp1  cosh[(1 + i ) y / δ v ]  iω 
= e  1 −  − p1 
dy 2 1+ i  ω K dx dx  cosh[(1 + i ) y0 / δ v ]  K 
(C.11)
Integration of Eq. (C.10) and (C.11) yields
1 δκ  ic dTm dp1 iωp1 
C1 ( y ) = e − (1+ i ) y / δ κ  p −  ⋅ (− δ κ )
2 (1 + i ) 2
  ω K dx dx K 
ic p dTm dp1 1 1
− ⋅ −2
ωK dx dx cosh[(1 + i ) y0 / δ v ] δ v − δ κ− 2
 cosh[(1 + i ) y / δ v ] sinh[(1 + i ) y / δ v ]  
⋅ +  , (C.12)
 δκ δv 
1 δκ (1+ i ) y / δ κ  ic p dTm dp1 iωp1 
C2 ( y ) = e  −  ⋅ (− δ κ )
2 (1 + i ) 2
 ωK dx dx K 
ic dTm dp1 1 1
− p ⋅ −2
ωK dx dx cosh[(1 + i ) y0 / δ v ] δ v − δ κ− 2
 cosh[(1 + i ) y / δ v ] sinh[(1 + i ) y / δ v ]  
⋅ −  . (C.13)
 δκ δv 
Substitution of Eq. (C.12) and (C.13) into Eq. (C.8) yields the particular solution
p1 1 dTm dp1 1 σ dTm dp1 cosh[(1 + i ) y / δ v ]
T1* = − − ,
ρ mc p ρ mω dx dx ρ mω 1 − σ dx dx cosh[(1 + i ) y0 / δ v ]
2 2

(C.14)
where σ = c p µ / K = δ / δ κ is the Prandtl number.
v
2 2

The summation of solutions of the homogenous equation and the particular


solution yields the final general solution
T1 = C1 ⋅ e(1+ i ) y / δ κ + C2 ⋅ e −(1+ i ) y / δ κ + T1*
220 Appendix

p1 1 dTm dp1
= C1 ⋅ e(1+ i ) y / δ κ + C2 ⋅ e − (1+ i ) y / δ κ + −
ρ m c p ρ mω 2 dx dx
1 σ
dTm dp1 cosh[(1 + i ) y / δ v ]
− . (C.15)
ρ mω 1 − σ dx dx cosh[(1 + i ) y0 / δ v ]
2

Next, the boundary conditions are applied to get the coefficients C1 and C 2 .
Boundary conditions:
∂T1
e) at y = 0 , because of the symmetry, =0
∂y
f) at y = y 0 , because of the solid wall, T1 = Tb1
By using boundary condition e, we obtain
C1 = C 2 . (C.16)
By using boundary condition f, the coefficients are obtained
1  
C1 =  Tb1 − p1 − 1 2 dTm dp1 1  . (C.17)
2 ⋅ cosh[(1 + i ) y0 / δ κ ]  ρ m c p ρ mω dx dx σ − 1 
Substitution of Eq. (C.16) and (C.17) into Eq. (C.15) yields the final solution of Eq.
(C.1)
cosh[(1 + i ) y / δ κ ]  p 1 dTm dp1 1  p
T1 = Tb1 − 1 − + 1

cosh[(1 + i ) y0 / δ κ ]  
ρ m c p ρ mω dx dx σ − 1  ρ m c p
2

1 dTm dp1 1 σ dTm dp1 cosh[(1 + i ) y / δ v ]


− − . (C.18)
ρ mω dx dx ρ mω 1 − σ dx dx cosh[(1 + i ) y0 / δ v ]
2 2

Now, the only unknown parameter is the temperature Tb1 at the surface where the
fluid and the solid wall have contact, i.e. y = y0 and y′ = l . As the heat flow into
the fluid and the solid wall at the boundary has the same amount but opposite in
direction. It means the following boundary condition is true
K (∂T1 / ∂y ) y = y = − K s (∂Ts1 / ∂y′) y ′ = l . (C.19)
0

From Eq. (2.1.24), it can be obtained that


∂Ts1 1+ i
= Tb1 ⋅ ⋅ tanh[(1 + i )l / δ s ] . (C.20)
∂y′ y ′=l
δs
From Eq. (C.18), it can be obtained that
∂T1  p 1 dTm dp1 1  1 + i
=  Tb1 − 1 − ⋅ ⋅ tanh[(1 + i ) y0 / δ κ ]
∂y  ρ c ρ ω 2
dx dx σ − 1  δ
y = y0  m p m  κ

1 dTm dp1 σ 1 + i
+ ⋅ ⋅ tanh[(1 + i ) y0 / δ v ] . (C.21)
ρ mω 2 dx dx σ − 1 δ v
Substitution of Eq. (C.20) and (C.21) into (C.19) yields
Appendix 221

−1
 tanh[(1 + i ) y0 / δ κ ] K s tanh[(1 + i )l / δ s ] 
Tb1 =  + 
 δκ K δs 
 p 1 dTm dp1 1  tanh[(1 + i ) y0 / δ κ ]
⋅  1 + ⋅
 ρ m c p ρ mω dx dx σ − 1  δκ
2

1 dTm dp1 σ tanh[(1 + i ) y0 / δ v ] 


− ⋅ . (C.22)
ρ mω 2 dx dx σ − 1 δv 
Substitution of Eq. (C.22) into (C.18) yields the final solution for the temperature
oscillation of the fluid layer:
p1 1  σ cosh[(1 + i ) y / δ v ]  dp1 dTm
T1 = − × 1 − 
ρ mc p ρ mω  (σ − 1)cosh[(1 + i ) y0 / δ v ]  dx dx
2

 p
− 1 +
(dp1 dx )(dTm / dx ) 1 + ε s f v  × cosh[(1 + i ) y / δ κ ] ,
 ρ m c p (σ − 1)ρ mω 2  f k  (1 + ε s )cosh[(1 + i )y0 / δ κ ] (C.23)

where the Rott’s functions are as defined in Eqs. (2.1.32), (2.2.33), and (2.1.34):
tanh[(1 + i ) y 0 / δν ]
fν = , (2.1.32)
(1 + i ) y 0 / δν
tanh[(1 + i ) y 0 / δ κ ]
fκ = , (2.1.33)
(1 + i ) y 0 / δ κ
Kρ m c p tanh[(1 + i ) y0 / δ κ ]
εs = .
K s ρ s cs tanh[(1 + i )l / δ s ]
(2.1.34)
222 Appendix

D: derivation of the time-averaged total energy


flow
The time-averaged product of two oscillatory quantities with the same frequency,
for example, a current I = I A cos(ω t + φI ) and voltage V = VA cos(ω t + φV ) , is
given by
1 τ 1 τ
∫ I A cos(ω t + φI )VA cos(ω t + φV )dt
τ ∫0
IV = IVdt =
τ 0

1
= I AVA cos(φI − φV ) . (D.1)
2
with τ = 2π / ω is the period of the oscillation.
If write the oscillatory current and voltage in complex notation, then, they are:
I = I Aei (ω t +φ I ) = I Aeiω t and V = VAei (ω t +φV ) = VAeiω t , respectively, where
I A = I Aeiφ I and VA = VAeiφV are the complex amplitudes.
The time-averaged value can be computed by
IV =
1
2
~
[
1
] [ 1
]
Re I A VA = Re I Aeiφ I ⋅ VAe − iφV = I AVA cos(φI − φV ) .
2 2
(D.2)

where the tilde denotes complex conjugation.


By using Eq. (D.2), the Eq. (2.1.71) can be rewritten as
E& 2 1 y0
∏ 2 0
[
= ∫ ρ mc p Re[T1u1 0 s ]
~ ] dy − ( y K + lK ) dTm .
dx
(D.3)

Rewrite Eq. (D.3) as


E& 2 1  y0
∏ 2 
 0 1 [
 0 s ]
~ dy − ( y K + lK ) dTm .
= Re ∫ ρ mc p T1u
dx
(D.4)

First, we focus on the term in the outer square bracket. For simplicity of writing,
this part is defined as
F=∫
y0

0
[ρ c T u~ ]dy .
m p 1 1 (D.5)

Substituting Eq. (2.1.30) for T1 and Eq. (2.1.20) for u 1 , Eq. (D.5) can be rewritten
as
y0 ~ 
− i dp cosh[(1 + i ) y / δ κ ] 
F=∫ 1
p1 1 − 
0 ωρ m dx  (1 + ε s )cosh[(1 + i ) y0 / δ κ ] 
  cosh[(1 + i ) y δ v ]  y 0 ic p ~ dT
dp1 dp
⋅ 1 − conj   dy + ∫ 1 m

  cosh [(1 + i ) y 0 δ ]
v 
0 ω3ρ
m dx dx dx
 σ cosh[(1 + i ) y / δ v ] 1 + ε s f v / fκ cosh[(1 + i ) y / δ κ ] 
⋅ 1 − + 
 (σ − 1)cosh[(1 + i ) y0 / δ v ] (σ − 1)(1 + ε s ) cosh[(1 + i ) y0 / δ κ ]
Appendix 223

  cosh[(1 + i ) y δ v ] 
⋅ 1 − conj   dy .
 cosh[(1 + i ) y0 δ v ] 
(D.6)

Here, conj ( function ) denotes taking complex conjugation of the function.
Eq. (D.6) consists of two integrations, denoting them as F1 and F2 , respectively,
for the sake of easy writing. Therefore, Eq. (D.4) can be written as
E& 2 1
= {Re[F1 ] + Re[F2 ]} − ( y0 K + lK s ) m .
dT
(D.7)
∏ 2 dx
For the first integration F1 , it can be obtained
 − iy0 dp ~
~  1 f v − fκ 
Re[F1 ] = Re 
~
1
p1 1 − f v +  .
 (D.8)
 ωρ m dx  1+ εs 1+σ 
After a long calculation and keeping in mind that only real terms are needed, the
term related to the second integration is
~ dT 1  ~ 1 + ε f / f f − ~
 ic p y0 dp1 dp f v 
Re[F2 ] = Re  3 1 m  fv +

s v κ
⋅ κ  (D.9)
 ω ρ m dx dx dx σ − 1  1+ εs 1 + σ 
Note that it is true for complex computation
Re[i (a + bi )] = Im[− (a + bi )] . (D.10)
Using Eq. (D.10) and substituting Eq. (D.8) and (D.9) into Eq. (D.7) yields
 dp ~
E& 2 ~  
1 f v − fκ
 − ( y0 K + lK s ) m
y0 ~ dT
= Im  1
p1 1 − f v +
∏ 2ωρ m  dx  
1+ εs 1+ σ
 dx
~ dT 1 ~
y c dp1 dp  ~ 1 + ε s f v / fκ fκ − f v 
+ 03 p 1 m
Im  f v + ⋅ . (D.11)
2ω ρ m dx dx dx 1 − σ  1+ εs 1+σ 
Therefore, the total energy flow along the stack is given by

E& 2 =
∏ y0  dp
~  ~
Im  1 p1 1 − fν −
(fκ − fν
~
) 

2ωρ m  dx  (1 + ε s )(1 + σ ) 
+
∏ y0 c p ~
dTm dp1 dp 1
~
× Im  fν +
( )
fκ − fν (1 + ε s fν / fκ ) 
~

2ω 3 ρ m (1 − σ ) dx dx dx  (1 + ε s )(1 + σ ) 
dT
− ∏( y 0 K + lK s ) m . (D.12)
dx
224 Appendix

E: derivation of the decoupling the sound field


into standing-wave and traveling-wave
components
The following is dedicated to the assumption that there exist standing-wave and
traveling-wave in the sound field inside a thermoacoustic device. Let’s start with
the wave equation, i.e. Eq. (2.1.57):
In the experiment, it is obvious that the stack plate material and the resonator tube
have a much larger thermal capacity compared with the working gas, i.e. ε s ≈ 0 .
The temperature span over the stack is small enough that the density and the Rott’s
functions f v , fκ do not have a dramatic change everywhere including the stack
zone.
Therefore, the wave equation can be simplified by realistic situation as
ρ m a 2 1 − f v d  d p1  a 2 f κ − fν dTm dp1
(1 + (γ − 1) f k )p1 +   − β =0.
ω 2 ρ m dx  dx  ω 2 (1 − σ ) dx dx
(E.1)
Using the assumption of ideal gas, i.e β = 1 / Tm , rewrite it as
d 2p1 f k − fv 1 dTm d p1 ω 2 1 + (γ − 1) f k
− ⋅ ⋅ ⋅ + 2 p1 = 0 .
dx 2 (1 − σ )(1 − f v ) Tm dx dx
(E.2)
a 1 − fv
In this experiment, the temperatures are all around ambient room temperature 300
K, and the temperature gradient is small.
fk − fv
The value of function is also small, roughly in the order of 1. So,
(1 − σ )(1 − f v )
the second term in Eq. (E.2) can be neglected. After all these simplification, the
wave equation now is written as
d 2p1 ω 2 1 + (γ − 1) f k
+ 2 p1 = 0 . (E.3)
dx 2 a 1 − fv
It is a second order of differential equation. The general solution of this differential
equation is:
p1 = A ⋅ ek 1 x + B ⋅ ek 2 x . (E.4)
k 1, 2 are the roots of the eigen equation y 2 + k 2 = 0 , where
ω 2 1 + (γ − 1) f k
k =
2
(E.5)
a2 1 − fv
Therefore, we obtain
Appendix 225

ω 1 + (γ − 1) f k
k 1, 2 = ± . (E.6)
a 1 − fv
Coefficients A and B depend on the boundary conditions.
Define a complex number
1 + (γ − 1) f k
χ + iξ = , (E.7)
1 − fv
and a real wave number
kr = ω / a . (E.8)
Rewrite k1, 2 as
k 1, 2 = ± k r (χ + iξ ) = ± (kr χ + ik rξ ) . (E.9)
Hence the acoustic pressure field is
p1 = A ⋅ e k r χx ⋅ eik r ξx + B ⋅ e − k r χx ⋅ e −ik r ξx . (E.10)
k r χx
Coefficients A and B depend only on the local boundary conditions. e and
− k r χx
e are real numbers depend on local conditions. The wave consists of incident
wave and reflected wave, if there is a change of acoustic impedance in the way of
k χx
wave propagating. The pressure reflection coefficient is the ratio of A ⋅ e r and
B ⋅ e − k r χx . Set a reference pressure in such a way that makes A ⋅ e k r χx have phase
of “0”, becoming a real number. Define two coefficients CL and C R as
CL = A ⋅ e k r χx and C R = B ⋅ e − k r χx (E.11)
Rewrite equation (E.10) as
p1 = CL ⋅ eik r ξx + C R ⋅ e − ik r ξx , (E.12)
Or
p1 = CL ⋅ (cos(kr χx) + i sin(kr χx) ) + C R ⋅ (cos(kr χx) − i sin(kr χx) ) . (E.13)
After some mathematical rearrangement, Eq. (E.13) can be written as
p1 = 2CL ⋅ cos(kr χx) + (C R − CL ) ⋅ e − ik r χx . (E.14)
The coefficient of the second term of Eq. (E.14), C R − CL , is a constant complex
number, writing it as
C R − C L = R A e iφ r , (E.15)
where RA and φr are amplitude and phase of the complex number C R − CL .
Substitution of Eq. (E.15) into (E.14), yields
p1 = 2CL ⋅ cos(kr χx) + RA ⋅ e − i ( k r χx −φ r ) . (E.16)
From equation (E.16), it is clear that the acoustic wave consists of two components
locally, standing-wave component 2C L ⋅ cos(k r χx) and a right-going traveling-
wave component RA ⋅ e − i ( k r χx −φ r ) .
226 Appendix

F: computation of loop section in a traveling-


wave system
Computation station A: The length of the feedback inertance tube is indicated as
L fb and the diameter as d fb . By means of the same method as did to obtain Eq.
(4.3.26) and Eq. (4.3.27), the pressure and volume velocity at computation station
A are found to be given by
4U1 fb ρ m a
p1 A = p1 fb cos(kL fb ) − i sin(kL fb ) , (F.1)
πd 2fb
πd 2fbp1 fb
U1 A = U1 fb cos(kL fb ) − i sin(kL fb ) . (F.2)
4ρma
If we use the concept of transfer matrix, it is helpful to understand the method in
this appendix. For the acoustic pressure and volume velocity at the two ends of any
component in the loop section can be related by a transfer matrix:
 p1   A11 A12   p1 
U  =  A A 22  U1  b
, (F.tm1)
 1  a  21
where the subscription, a and b, denote the two ends of a component.
For easy writing, here the transfer matrix is denoted as:

[TM ]a −b = 
A11 A12 
A 22 
. (F.tm2)
 A 21
If write (F.1) and (F.2) in the form of transfer matrix, the transfer matrix is given as:
 4ρ a 
 cos(kL fb ) − i m2 sin(kL fb )
πd fb
[TM ]A− fb = . (F.tm3)
 πd fb
2

− i 4 ρ a sin(kL fb ) cos(kL fb ) 
 m 

Computation station B: Between A and B, is a volume having a character of a


compliance. Normally, the geometry depends on specific machine. If the geometry
parameters are known, the pressure and volume velocity at B can be obtained by
the same method used for proceeding results by splitting the volume into
elementary computation cells. Here, for simplicity, this compliance volume is
considered as a tube segment of length Lcpl with a large diameter d cpl . Its diameter
is larger than those of feedback tube and the ambient heat exchanger. Repeating the
same calculation, the pressure and volume velocity at B are
Appendix 227

4U1 A ρ m a
p1B = p1 A cos(kLcpl ) − i sin(kLcpl ) , (F.3)
πd cpl
2

πd cpl
2
p1 A
U1B = U1 A cos(kLcpl ) − i sin(kLcpl ) . (F.4)
4ρma
The transfer matrix is given as:
 4ρ a 
 cos(kLcpl ) − i m2 sin(kLcpl )
πd cpl
[TM ]B − A = . (F.tm4)
 πd cpl
2

− i 4 ρ a sin(kLcpl ) cos(kLcpl ) 
 m 

Computation station C: The assumption of a zero-temperature gradient does hold


for the heat exchangers. Therefore, Eqs. (4.3.13) to (4.3.16) can be used to the
computation of heat exchangers. In practice, to get good thermal contact with the
gas, the hydraulic radius of the heat exchangers, especially the ambient heat
changer, is of the order of working gas’ thermal penetration depth, i.e. f v and
fκ having order of 1. Thus, the simplifications used for big channel are not valid
any more. The ambient heat exchanger has a porosity ψ aHX , diameter d aHX , and
length LaHX . Start with Eq. (4.3.13) and (4.3.16), applying the boundary conditions
and using the same method, the pressure and volume velocity at C can be obtained
4U1Bωρ m
p1C = p1B cos(k aHX LaHX ) − i sin(k aHX LaHX ) ,
(1 − f v )aHX k aHXψ aHX πd aHX
2 (F.5)

p1B (1 − f v )aHX k aHXψ aHX πd aHX


2
U1C = U1B cos(k aHX LaHX ) − i sin(k aHX LaHX ) ,
4ωρ m
(F.6)
where the complex wave number is evaluated at temperature of ambient heat
exchanger and the Rott’s functions computed with hydraulic radius of the ambient
heat exchanger
ω 2 1 + (γ − 1) fκ /(1 + ε s )
k 2
= ⋅ , (F.7)
1 − fv
aHX
a2
and the hydraulic radius is
ψ aHX πd aHX
2
rh − aHX = . (F.8)
4Π aHX
In practical situations, the lengths of the heat exchangers are short compared with
the complete system. Here, for simplicity, it is assumed that all the heat exchangers
are ideal ones (no blockage, porosity=100%). Therefore, we have
228 Appendix

p1C = p1B and U1C = U1B . (F.9)


The transfer matrix is given as:

[TM ]C − B = 
1 0
 (F.tm5)
0 1 

Regenerator C-D: Between C and D is the regenerator. There is a strong


temperature gradient along the regenerator. The assumption of zero-temperature
gradient is not valid for the regenerator. Therefore, Eqs. (4.3.13) to (4.3.16) are not
valid for regenerator. In this work, a distributed model for the regenerator proposed
by Backhaus and Swift [34] is employed.

U1,up U1,1 U1, 2 U1, N U1,down


p1,1 p1,2 p1, N
p1,up ∆R1 ∆R2 ∆RN p1,down
∆C ∆C ∆C Tdn−st
Tup − st U1, 2 ∆T2 U1, N ∆TN
U1,1∆T1
Tm,1 Tm , 2 Tm , N

Figure F.1: Distributed model of the regenerator.

In this distributed model, the regenerator of length Lreg is split into N = Lreg / ∆x
segments. Each segment is of length ∆x and has a temperature span of ∆Tn , as
shown in Fig. F.1. With an ideal gas as the working fluid, the compliance of each
segment of the regenerator is the isothermal form
∆C = C0 ∆x / Lreg , (F.10)
where C0 = ψ reg Areg Lreg / pm . ψ reg and Areg are the volume porosity and cross-
sectional area of the regenerator. Tm is the local mean (average) temperature in the
regenerator. Due to the temperature distribution along the regenerator and the
inside compliance, the volumetric velocity changes across each regenerator
segment, i.e.
U1, n +1 − U1, n = U1, n ∆Tn / Tm , n − iω∆Cp1, n . (F.11)
Dividing both sides of the equation by ∆x and letting ∆x → 0 yields the
differential equation for U1 :
dU1 U1 dTm C
= − iω 0 p1 . (F.12)
dx Tm dx Lreg
Appendix 229

The fluid resistance of each segment of the regenerator is given by



R ∆x  T 
∆Rn = 0 ⋅  m , n  , (F.13)
Lreg  T0 
where
R0 ≈ 3µ0 Lreg /(ψ reg Areg rh2− reg ) (F.14)
is the low-Reynolds-number-limit flow resistance of the regenerator when its entire
length is at ambient temperature T0 . Over a wide range in temperature, the
viscosity of gases can be described as µ = µ0 (T / T0 ) µ , where bµ is a constant
b

depending on the gas. The viscosity temperature dependence is counted in the


resistance by the factor of Tm , n / T0( )bµ
The pressure drop across each regenerator
segment is mainly caused by the fluid resistance, i.e.
p1, n − p1, n −1 = −∆Rn U1, n . (F.15)
Similarly, the differential equation for the pressure is obtained by dividing through
with ∆x and letting ∆x → 0 .

dp1 R  Tm 
=− 0   U1 . (F.16)
dx Lreg  T0 
Assume that the temperature has a linear distribution axially along the regenerator,
which is a good approximation to the real case. Therefore, the temperature is given
by
Tdn − st − Tup − st  1/τ − 1 
Tm ( x) = Tup − st + x = Tup − st 1 + x , (F.17)
Lreg  Lreg 
where the temperature ratio
τ = Tup − st / Tdn − st . (F.18)
In this cooler mode, Tup − st is the ambient temperature T0 and the Tdn − st is the given
cold end temperature TC . In the mode of engine, Tup − st is the ambient
temperature T0 and the Tdn − st is the given hot end temperature TH .
Under typical operation conditions, the pressure drop across the regenerator is
small, less than 10% [34], i.e. ∆p1,regen ≈ 0.1 p1,up . Thus, p1 ( x ) = p1,up is assumed
and substituted into Eq. (F.12). The integration of Eq. (F.12) yields
 1/τ − 1  iωC0p1, up  1 / τ − 1 
U1 = 1 + x  U1,up − ln1 + x  . (F.19)
 L  1 / τ − 1  L 
 reg    reg 
Now Eq. (F.19) is substituted into Eq. (F.16) and integrated it along the regenerator.
The linear temperature distribution is also assumed.
230 Appendix

The pressure drop is given by:


∆p1reg = p1C − p1D = U1, up R0 f (τ , bµ ) + iωC0 R0p1,up g (τ , bµ ) , (F.20)
( ) (
where the functions f τ , bµ and g τ , bµ are defined as:)
b +2
(1 / τ ) µ − 1
f (τ , bµ ) = , (F.21)
(1 / τ − 1)(bµ + 2)
 (1 / τ )bµ + 2 − 1 (1 / τ )bµ + 2 ln (1 / τ ) 
g (τ , bµ ) =
1
 − . (F.22)
(1 / τ − 1) 2  (bµ + 2) 2 bµ + 2 
Substitution of x = Lreg , U1,up = U1C and p1, up = p1C in Eq. (F.19), the volume
velocity at the cold end of the regenerator is obtained in terms of the inlet
conditions,
U1C iωC0p1C ln τ
U1D = + . (F.23)
τ 1 −τ
Substitution of U1,up = U1C and p1, up = p1C into Eq. (F.20) yields an expression for
the pressure in terms of inlet pressure and volume velocity:
( )
p1D = p1C 1 − iωC0 R0 g (τ , bµ ) − U1C R0 f (τ , bµ ) . (F.24)
The transfer matrix is given as:
1 − iωC0 R0 g (τ , bµ ) − R0 f (τ , bµ )
[TM ]D −C =  iωC0 lnτ 1  (F.tm6)
 
 1−τ τ 

Computation station E: Notice that we assume all the heat exchangers are ideal
ones in the above part. Therefore, the pressure and volume velocity across the cold
end heat exchanger are considered unchanged:
p1E = p1D , U1E = U1D , (F.25)
The transfer matrix is given as:

[TM ]E − D = 
1 0
 (F.tm7)
0 1 

Computation station E-Tee: The length from the surface E to the joint tee is
indicated as Ltb and the diameter as d tb . Using the same method as before and
considering Eq.(F.25), the pressure and volume velocity at the tee branch
connected to the cold-end heat exchanger are given by:
4U1D ρ m a
p1out = p1D cos(kLtb ) − i sin(kLtb ) , (F.26)
πdtb2
Appendix 231

πd tb2 p1D
U1out = U1D cos(kLtb ) − i sin(kLtb ) . (F.27)
4ρm a
The transfer matrix is given as:
 4ρ a 
 cos(kLtb ) − i m2 sin(kLtb )
πdtb
[TM ]out − E = [TM ]out − D =  (F.tm8)
− i πd tb sin(kL )
2
cos(kLtb ) 
 4 ρ m a tb


Direct relation between p1out , U1out and p1 fb , U1 fb :


Now the relations of acoustic pressures and volume velocities between all the
subsequent computation stations from A to E are obtained. It can be seen that all
the acoustic pressures and volume velocities can be expressed in terms of
p1 fb and U1 fb . The following analysis to find the relation between p1 fb , U1 fb and
p1out , U1out is carried out. The end result is looking for the impedance that the
loudspeaker “sees” when looking into the resonator.
Substitution of Eq. (F.1) and (F.2) into Eq. (F.3) and (F.4) yields
 d 2fb 
p1B = p1 fb cos(kL fb ) cos(kLcpl ) − 2 sin(kL fb ) sin(kLcpl )
 d cpl 
4ρ a
π
[
− i m sin(kL fb ) cos(kLcpl ) / d 2fb + cos(kL fb ) sin(kLcpl ) / d cpl
2
]
U1 fb , (F.28)

 d2 
U1B = U1 fb cos(kL fb ) cos(kLcpl ) − cpl
2
sin(kL fb ) sin(kLcpl )
 d fb 
π
−i
4ρm a
[d 2fb sin(kL fb ) cos(kLcpl ) + d cpl
2
cos(kL fb ) sin(kLcpl )]p1 fb . (F.29)

By substituting Eq. (F.9) into Eq. (F.23) and (F.24), the acoustic pressure and
volume velocity are given by
U1B iωC0 ln τ
U1D = + p1B , (F.30)
τ 1−τ
p1D = p1B (1 − iωC0 R0 g (τ , bµ ) ) − U1B R0 f (τ , bµ ) . (F.31)
Substitution of Eq. (F.28) and (F.29) into Eq. (F.30) and (F.31) yields
{[
U1D = U1 fb cos(kL fb ) cos(kLcpl ) / τ − d cpl
2
sin(kL fb ) sin(kLcpl ) /(τd 2fb ) ]
4 ρ m aωC0 ln τ
+
π (1 − τ )
[ 2 
sin(kL fb ) cos(kLcpl ) / d 2fb + cos(kL fb ) sin(kLcpl ) / d cpl  ]

 iωC0 ln τ
+ p1 fb  [
cos(kL fb ) cos(kLcpl ) − d 2fb sin(kL fb ) sin(kLcpl ) / d cpl
2
]
 1 −τ
232 Appendix



4 ρ m aτ
[d 2fb sin(kL fb ) cos(kLcpl ) + dcpl
2
] 
cos(kL fb ) sin(kLcpl )  ,
(F.32)

p = p {(1 − iωC R g (τ , b ) )[cos(kL ) cos(kL ) − d sin(kL ) sin(kL )
1D 1 fb 0 0 µ fb cpl
2
fb fb cpl

πR f (τ , b )
/ d ]+ i
2
[d sin(kL ) cos(kL ) + d cos(kL ) sin(kL )]}
0 µ 2 2

4ρ a
cpl fb fb cpl cpl fb cpl
m

-U {R f (τ , b )[cos(kL ) cos(kL ) − d sin(kL ) sin( kL ) / d ]


1 fb 0 µ fb cpl
2
cpl fb cpl
2
fb

4ρ a
+ i (1 − iωC R g (τ , b ) )
0
π
[sin(kL ) cos(kL ) / d
0 µ
m
fb cpl
2
fb

+ cos(kL ) sin(kL ) / d ]}.


fb cpl
2
cpl (F.33)
To make U1D and p1D easy to write, they are rewritten as
p1D = D1 ⋅ p1 fb + D 2 ⋅ U1 fb , (F.34)
U1D = D3 ⋅ p1 fb + D 4 ⋅ U1 fb . (F.35)
The complex coefficients from D1 to D4 only depend on the geometry of the
refrigerator and operational conditions. They are given by
[
D1 = (1 − iωC0 R0 g (τ , bµ ) ) cos(kL fb ) cos(kLcpl ) − d 2fb sin( kL fb ) sin(kLcpl ) / d cpl
2
]
πd 2fb R0 f (τ , bµ )
+i
4ρm a
[sin(kL fb ) cos(kLcpl ) + dcpl2 cos(kL fb ) sin(kLcpl ) / d 2fb ] (F.36a)

D2 = − R0 f (τ , bµ )[cos(kL fb ) cos(kLcpl ) − d cpl 2


sin(kL fb ) sin(kLcpl ) / d 2fb ]
4ρ a
− i (1 − iωC0 R0 g (τ , bµ ) ) m2 [sin(kL fb ) cos(kLcpl )
πd fb
+ d 2fb cos(kL fb ) sin(kLcpl ) / d cpl
2
] (F.36b)
iωC0 ln τ
D3 =
1 −τ
[
cos(kL fb ) cos(kLcpl ) − d 2fb sin(kL fb ) sin(kLcpl ) / d cpl
2
]
iπd fb
[ ]
2

− sin(kL fb ) cos(kLcpl ) + d cpl


2
cos(kL fb ) sin(kLcpl ) / d 2fb (F.36c)
4 ρ m aτ
[
D4 = cos(kL fb ) cos(kLcpl ) − dcpl
2
sin(kL fb ) sin(kLcpl ) / d 2fb / τ ]
4 ρ m aωC0 lnτ
+
πd fb (1 − τ )
2
[
sin(kL fb ) cos(kLcpl ) + d 2fb cos(kL fb ) sin(kLcpl ) / d cpl
2
. ] (F.36d)

For the simplicity of writing, four real functions are defined, which only depend on
the geometrical configurations:
θ1 = cos(kL fb ) cos(kLcpl ) − d 2fb sin(kL fb ) sin(kLcpl ) / d cpl
2
(F.37a)
θ 2 = cos(kL fb ) cos(kLcpl ) − d sin(kL fb ) sin(kLcpl ) / d 2
cpl
2
fb (F.37b)
Appendix 233

θ3 = sin(kL fb ) cos(kLcpl ) + d cpl


2
cos(kL fb ) sin(kLcpl ) / d 2fb (F.37c)
θ 4 = sin(kL fb ) cos(kLcpl ) + d cos(kL fb ) sin(kLcpl ) / d .
2
fb
2
cpl (F.37d)
Therefore, the complex coefficients of D can be rewritten as
πd 2fb R0 f (τ , bµ )
D1 = (1 − iωC0 R0 g (τ , bµ ) )θ1 + i θ3 (F.38a)
4ρma
4ρ a
D2 = − R0 f (τ , bµ )θ 2 − i (1 − iωC0 R0 g (τ , bµ ) ) m2 θ 4 (F.38b)
πd fb
iωC0 ln τ iπd 2fb
D3 = θ1 − θ3 (F.38c)
1−τ 4 ρ m aτ
4 ρ aωC ln τ
D4 = θ 2 / τ + m 2 0 θ . (F.38d)
πd fb (1 − τ ) 4
Substituting Eq. (F.34) and (F.35) into Eq. (F.26) and (F.27), the acoustic pressure
p1out and volume velocity U1out can be written in terms of p1 fb and U1 fb :
 4 ρ aD 
p1out = p1 fb D1 cos(kLtb ) − i m 2 3 sin(kLtb )
 πd tb 
 4 ρ aD 
+ U1 fb D 2 cos(kLtb ) − i m 2 4 sin( kLtb ) , (4.3.28)
 πd tb 
 πdtb2 D1 
U1out = p1 fb D3 cos(kLtb ) − i sin(kLtb )
 4ρma 
 πd D2

+ U1 fb D4 cos(kLtb ) − i tb 2 sin(kLtb ) . (4.3.29)
 4ρma 
If we use the form of transfer matrix, computation stations from A to B in the form
of Eq.(F.tm1) are given as:
 p1   p1 
U  = [TM ]A − fb U  , (F.tm8)
 1 A  1  fb
 p1   p1 
U  = [TM ]B − A U  . (F.tm9)
 1B  1 A
By substituting Eq.(F.tm8) into (F.tm9), the transfer matrix between B and fb is
obtained as:
[TM ]B − fb = [TM ]B − A [TM ]A− fb . (F.tm10)
Repeating this computation for stations C, D, E to “out” (the tee branch connected
to the cold-end heat exchanger), the transfer matrix between “out” and “fb” is
obtained as:
234 Appendix

[TM ]out − fb = [TM ]out − E [TM ]E − D [TM ]D −C [TM ]C − B [TM ]B − A [TM ]A− fb .
(F.tm11)
If write Eqs.(4.3.28) and (4.3.29) in the form of transfer matrix, the transfer matrix
is given as:
[TM ]out − fb =
 4 ρ m aD 3 4 ρ m aD 4 
D1 cos( kLtb ) − i πd 2 sin( kLtb ) D2 cos( kLtb ) − i πd 2 sin( kLtb )
 tb tb
.
π 2
 D cos( kL ) − i tb 1 sin( kL ) π 2
sin( kLtb ) 
d D d D
D 4 cos( kLtb ) − i tb 2
 3 tb
4ρma
tb
4ρm a 
(F.tm12)
Therefore, the direct relation between “out” and “fb” in the form of transfer matrix
is given as:
 p1   p1 
U  = [TM ]out − fb U  , (F.tm13)
 1  out  1  fb
Appendix 235

G: transmission of acoustic impedance of a


uniform pipe
Next, the transmission of acoustic impedance of a pipe is derived. Because the set-
up we use is connected to the looped tube via a long straight pipe element. Two
random positions in a tube are indicated by xb and xc in the direction of acoustic
wave propagation, showed in Fig. G.1. By using the Eq. (4.3.19) and (4.3.20), the
acoustic impedance is given by
p1 ωρ m C2 ⋅ e- ikx + C1 ⋅ eikx
Z= =
U1 (1 − fν ) Ak C2 ⋅ e- ikx − C1 ⋅ eikx
(G.1)

o x xb xc

Figure G.1: Two random positions along a tube.

Splitting the exponential factors into cosine and sine and combining terms, Eq.
(G.1) becomes
C2 − C1
cos(kx) − i sin(kx)
ωρ m C2 + C1
Z= .
(1 − fν )Ak C2 − C1
cos(kx) − i sin(kx)
(G.2)

C2 + C1
Eq. (G.2) is the acoustic impedance at x. Replacing x with xb and xc . The acoustic
impedances of them are respectively
C 2 − C1
cos(kxb ) − i sin(kxb )
ωρ m C2 + C1
Zb = ,
(1 − fν )Ak
C2 − C1
cos(kxb ) − i sin(kxb )
(G.3)

C2 + C1
C − C1
cos(kxc ) − i 2 sin(kxc )
ωρ m C2 + C1
Zc = .
(1 − fν )Ak C2 − C1 cos(kx ) − i sin(kx ) (G.4)

C2 + C1
c c

Now, the impedance Z con is defined as:


236 Appendix

ωρ m
Z con = .
(1 − fν )Ak (G.5)

Eliminating (C2 − C1 ) /(C2 + C1 ) , and using (G.3) (G.4), one finds the relation of
acoustic impedance between these two positions:
Zb
cos[k ( x c − xb )] − i sin[k ( x c − xb )]
Zc Z con
= (G.6)
Z con Z
cos[k ( x c − xb )] − i b sin[k ( x c − xb )]
Z con
Or
Zc
cos[k ( x c − xb )] + i sin[k ( x c − xb )]
Zb Z con
= . (G.7)
Z con Z
cos[k ( x c − xb )] + i c sin[k ( x c − xb )]
Z con
This is the transmission relation for the acoustic impedance.
For open channel tube, i.e. the hydraulic radius rh >> δ v and δ κ can be sustained,
thus Rott’s functions f v ≈ 0 and fκ ≈ 0 , Eq. (G.5) can be reduced to
aρ m
Z con = . (G.8)
A
In that case, the transmission equations become much simpler,
ρm a
Z b cos[k ( x c − xb )] − i sin[k ( x c − xb )]
ρma A
Zc = (G.9)
A ρ m a cos[k ( x − x )] − iZ sin[k ( x − x )]
c b b c b
A
Or
ρma
Z c cos[k ( x c − xb )] + i sin[k ( x c − xb )]
ρma A
Zb = . (G.10)
A ρ m a cos[k ( x − x )] + iZ sin[k ( x − x )]
c b c c b
A
Appendix 237

H: Fortran code for computation of traveling-


wave engine
* simple program for traveling-wave machine
* edited on 13 Feb., 2009
*-------------------------------------------------------------------
Real R,gama,b_miu,pi,f,round_f,Pm,P_in, T_cold,T_hot,density,
& conduct_fluid,conduct_solid,Cp_fluid,sound_speed,miu,Pr,
& k,L_fb,d_fb,L_cpl,d_cpl,porosity,d_reg,L_reg,r_h,area_gas,
& area_solid,L_tb,d_tb,C0,R0,tau,f_reg,g_reg,sita1,sita2,
& sita3,sita4,pen_f_d,v_pen_fluid,ratio_v,ratio_k,term1,F3,
& term2,term3,F2,Q_c,U_mu,U_angle,U_mu1,U_angle1,U_mui,U_anglei
& ,power,inertance,compliance,U_mu2,U_angle2,U_mu3,U_angle3,
& U_mu4,U_angle4
complex i,f_v,t_v,f_k,t_k,D1,D2,D3,D4,coef_i1,coef_i2,p_D,U_D,
& coef_i3,Z_fb,Z_1,Z_input,U_input,U_fb,Z_i,U_i,U_1fb,U_1c,
& p_out,U_out,U_1out
open(21, file='out_dw.dat')
R=8.3145/(4.003e-3)
gama=5.0/3.0
b_miu=0.68
pi=3.1416
i=cmplx(0.0,1.0)
f=84.12
Pm=3103000.0
p_in=2.723e5
T_hot=300.0
T_cold=900.0
density=Pm/(R*T_hot)
conduct_fluid=0.156084
Cp_fluid=5.19312*1000.0
sound_speed=1019.56
miu=1.9937e-5
Pr=Cp_fluid*miu/conduct_fluid
conduct_solid=40.0
round_f=2.0*3.14*f
k=round_f/sound_speed
L_fb=(10.2+25.6+20.9)*0.01
d_fb=7.8*0.01
238 Appendix

L_cpl=0.348
d_cpl=10.2*0.01
porosity=0.72
d_reg=8.89*0.01
L_reg=7.3*0.01
r_h=42.0e-6
area_gas=pi*d_reg**2*porosity/4.0
area_solid=pi*d_reg**2*(1.0-porosity)/4.0
L_tb=(24.0+1.0+7.0)*0.01
d_tb=8.89*0.01
C0=porosity*pi*d_reg**2*L_reg/4.0/Pm
R0=4.5*miu*L_reg/(porosity*0.25*pi*d_reg**2*r_h**2)
tau=T_hot/T_cold
f_reg=((1.0/tau)**(b_miu+2.0)-1.0)/(1.0/tau-1.0)/(b_miu+2.0)
g_reg=(((1.0/tau)**(b_miu+2.0)-1.0)/(b_miu+2.0)**2-(1.0/tau)**
& (b_miu+2.0)*log(1.0/tau)/(b_miu+2.0))/(1.0/tau-1.0)**2
sita1=cos(k*L_fb)*cos(k*L_cpl)-
& d_fb**2*sin(k*L_fb)*sin(k*L_cpl)/d_cpl**2
sita2=cos(k*L_fb)*cos(k*L_cpl)-
& d_cpl**2*sin(k*L_fb)*sin(k*L_cpl)/d_fb**2
sita3=sin(k*L_fb)*cos(k*L_cpl)+
& d_cpl**2*cos(k*L_fb)*sin(k*L_cpl)/d_fb**2
sita4=sin(k*L_fb)*cos(k*L_cpl)+
& d_fb**2*cos(k*L_fb)*sin(k*L_cpl)/d_cpl**2
D1=(1.0-i*round_f*C0*R0*g_reg)*sita1+i*pi*d_fb**2*R0*f_reg*sita3
& /(4.0*density*sound_speed)
D2=-R0*f_reg*sita2-i*(1.0-i*round_f*C0*R0*g_reg)*sita4*
& 4.0*density*sound_speed/(pi*d_fb**2)
D3=i*round_f*C0*log(tau)*sita1/(1.0-tau)-i*pi*d_fb**2*sita3/
& (4.0*density*sound_speed*tau)
D4=sita2/tau+4.0*density*sound_speed*round_f*C0*log(tau)*sita4/
& (pi*d_fb**2*(1.0-tau))

Z_fb=(D2*cos(k*L_tb)-i*4.0*density*sound_speed*D4*sin(k*L_tb)/
& (pi*d_tb**2))/(1.0-D1*cos(k*L_tb)+i*4.0*density*sound_speed*
& D3*sin(k*L_tb)/(pi*d_tb**2))

Z_1=pi*d_tb**2*D2/(4.0*density*sound_speed)+
& 4.0*density*sound_speed*D3/(pi*d_tb**2)
Appendix 239

Z_input=1.0/((D4*cos(k*L_tb)-i*pi*d_tb**2*D2*sin(k*L_tb)/
& (4.0*density*sound_speed)-1.0)/Z_fb+D3*cos(k*L_tb)-i*pi*
& d_tb**2*D1*sin(k*L_tb)/(4.0*density*sound_speed))

F3=real(D1*conjg(D3))+real(D2*conjg(D4))/(abs(Z_fb))**2
& +real(D2*conjg(D3)/Z_fb+D1*conjg(D4)/conjg(Z_fb))

U_input=p_in/Z_input
U_mu=cabs(U_input)
U_angle=180.0*atan(aimag(U_input)/real(U_input))/pi

U_fb=p_in/Z_fb
U_mu1=cabs(U_fb)
U_angle1=180.0*atan(aimag(U_fb)/real(U_fb))/pi
p_D=D1*p_in+D2*U_fb
U_D=D3*p_in+D4*U_fb
p_out=p_in*(D1*cos(k*L_tb)-i*4.0*density*sound_speed*D3*
& sin(k*L_tb)/(pi*d_tb**2))+U_fb*(D2*cos(k*L_tb)-i*4.0
& *density*sound_speed*D4*sin(k*L_tb)/(pi*d_tb**2))
U_out=p_in*(D3*cos(k*L_tb)-i*pi*d_tb**2*D1*sin(k*L_tb)/(4.0*
& density*sound_speed))+U_fb*(D4*cos(k*L_tb)-i*pi*d_tb**2*D2*
& sin(k*L_tb)/(4.0*density*sound_speed))

U_mu3=cabs(U_out)
U_angle3=180.0*atan(aimag(U_out)/real(U_out))/pi

pen_f_d=sqrt(2.0*conduct_fluid/
& (density*Cp_fluid*round_f))

v_pen_fluid=sqrt(2.0*miu/(density*round_f))
ratio_v=r_h/v_pen_fluid
call tanh(ratio_v,t_v)
f_v=t_v/(cmplx(1.0,1.0)*cmplx(ratio_v,0.0))

ratio_k=r_h/pen_f_d
call tanh(ratio_k,t_k)
f_k=t_k/(cmplx(1.0,1.0)*cmplx(ratio_k,0.0))
240 Appendix

coef_i3=conjg(f_v)+(f_k-conjg(f_v))/(1.0+Pr)

coef_i2=1.0-conjg(f_v)-(f_k-conjg(f_v))/(1.0+Pr)
coef_i1=(sita1*sita2/conjg(Z_fb)+sita3*sita4/Z_fb)*coef_i2

term1=area_gas*R0*aimag(coef_i1)/(2.0*round_f*density*L_reg)
term2=area_gas*R0*(pi*d_fb**2*sita1*sita3/(4.0*density*
& sound_speed)+4.0*density*sound_speed*sita2*sita4/
& pi/(d_fb*cabs(Z_fb))**2)*real(coef_i2)/
& (2.0*round_f*density*L_reg)
term3=area_gas*Cp_fluid*T_hot*R0**2*(1.0/tau-1.0)/L_reg**3
& /(2.0*round_f**3*density)/(1.0-Pr)*aimag(coef_i3)
& *((sita2/cabs(Z_fb))**2+(pi*d_fb**2*sita3/(4.0*
& density*sound_speed))**2+pi*d_fb**2*sita2*sita3
& *aimag(Z_fb)/(cabs(Z_fb))**2/(2.0*density*sound_speed))

Q_c=(abs(p_in))**2*(F3/2.0-F2)
& +(area_gas*conduct_fluid+area_solid*conduct_solid)
& *(1.0-1.0/tau)*T_hot/L_reg

Z_i=cmplx(1.09574e6,2.074e6)
U_i=p_in/Z_i
U_mui=cabs(U_i)
U_anglei=180.0*atan(aimag(U_i)/real(U_i))/pi

inertance=0.57e3
compliance=56.87e-11
U_1c=round_f**2*inertance*compliance/R0*p_in/
& (1.0+i*round_f*inertance/R0)
U_1fb=-i*round_f*compliance*p_in-U_1c
U_mu2=cabs(U_1fb)
U_angle2=180.0*atan(aimag(U_1fb)/real(U_1fb))/pi

U_1out=tau*U_1c
U_mu4=cabs(U_1out)
U_angle4=180.0*atan(aimag(U_1out)/real(U_1out))/pi

power=0.5*real(p_in*conjg(U_1c))*3.0
Appendix 241

write(21,*) 'DeltaE',' Qh=',3037.000,


& 'U_input= ',0.2870,' ',-89.452, ' U_fb= ',0.2029,
& ' ',86.355,' U_out= ',8.6321e-2,' ',-77.571
write(21,*) 'yan model',' Qh= ',Q_c, 'U_input='
& ,U_mu,' ',U_angle,' U_fb=',U_mu1,' ',U_angle1,
& ' U_out=',
& U_mu3,' ',U_angle3,' p_out=', p_out
write(21,*) 'swift smplified lump model', ' Qh=', power,
& 'U_input= ',U_mui,' ',U_anglei,' U_fb=',U_mu2,' ',
& U_angle2,' U_out=',U_mu4,' ',U_angle4
end
*==========================================================
* Subroutine tanh is used for calculate function tanh((1+i)y0/peneration depth)
* edit on 12 April, 2006
*==========================================================
subroutine tanh(alfa,f_tanh)
real alfa,r_part,i_part,B,C,D,E
complex f_tanh
B=(exp(alfa)-exp(-1.0*alfa))*cos(alfa)
C=(exp(alfa)+exp(-1.0*alfa))*sin(alfa)
D=(exp(alfa)+exp(-1.0*alfa))*cos(alfa)
E=(exp(alfa)-exp(-1.0*alfa))*sin(alfa)

r_part=(B*D+C*E)/(D**2+E**2)
i_part=(C*D-B*E)/(D**2+E**2)

f_tanh=cmplx(r_part,i_part)

end
242 Appendix

I: The design of the ambient heat exchanger in


the traveling-wave refrigerator
By the conclusion in reference [74], heat exchanger with ξhx/Lhx in the range of 3 to
8 can be thermally effective as a source or sink thermoacoustic heat transport if
y0/δκ is in the range of 0.75 to 0.5. The peak displacement amplitude of a gas parcel
in the heat exchanger is given by ξhx, and the length of the heat exchanger is given
by Lhx. The half separation between adjacent fins is y0 and δκ is the gas thermal
penetration depth. Hofler also pointed out that the acoustic loss or dissipation of the
heat exchanger is relatively less for y0/δκ=0.5. First, the gas thermal penetration
depth δκ is calculated for the designed working condition. By the definition, the
fluid’s thermal penetration depth is given as
2K
δκ = . (I.1)
ρ mc p ⋅ 2πf
Under the designed working condition: helium at mean pressure of 11 bar,
temperature of 300 K, and operating frequency of 320 Hz, the related gas
properties are given:
К=0.1567 W/(m·K), ρm=1.7571 kg/m³, and Cp=5.1925 kJ/(kg·K)
Substitution of the gas parameters into Eq. (I.1) gives the thermal penetration depth:
δκ =1.31×10-4 m. Considering the criterion y0/δκ=0.5, the separation between
adjacent fins is designed as:
2y0=δκ . (I.2)
The interfin distance should be 0.131 mm. By taking capability of manufacturing
into account, the distance between adjacent fins is designed as 0.17 mm.
By numerical simulation, the volume velocity in the ambient heat exchanger is
U1=5.95×10-3 m³/s. By using the relation between volume velocity and peak
displacement amplitude of a gas parcel
U1
ξ hx = , (I.3)
Ahx ⋅ 2πf
where Ahx is the cross sectional area of the heat exchanger.
The substitution of U1 and Ahx into (I.3) gives ξhx=4.275 mm. By the criterion
“ξhx/Lhx in the range of 3 to 8”, the length of heat exchanger should be in the range
of 0.534 mm to 1.425 mm. Further consideration on the criterion that small
temperature difference between the top and the root of a fin shows that such small
length Lhx will result in a large temperature span over a sigle fin.
A fin of the heat exchanger is shown in Fig. I.1. The length of a fin is lhx, and the
temperature of the root of a fin is Troot.
Appendix 243

y0 δfin
y0
Troot
Lhx
lfin

Figure I.1: Two adjacent fins of the heat exchanger.


The total amount of heat exchanged at the ambient heat exchanger is indicated as Φ.
The heat of unit cross-sectional area of the heat exchanger is given by:
Φ 4Φ
φ= = 2 . (I.4)
Ahx πd hx
The heat taken away by each fin is obtained as:
( )
φ1 = φ ⋅ 2 y0 + δ fin ⋅l fin . (I.5)

dx

Lhx
lfin
0 x
dx heat

Troot qx qx+dx δfin


x

Figure I.2: A fin and the cross-sectional view with an element for energy analysis.

An element in a fin, shown in Fig.I.2, is analyzed and the heat loaded from the
contact surface with gas is given by:
244 Appendix

φ1 φ1
⋅ dx ⋅ Lhx = ⋅ dx (I.6)
l fin ⋅ Lhx l fin
Considering Eq. (I.2), the heat flows in the element of the fin, which is dashed out
in Fig. I.2, follow:
φ1
− qx x ⋅ δ fin ⋅ Lhx + qx x + dx ⋅ δ fin ⋅ Lhx + ⋅ dx = 0 , (I.7)
l fin
where the qx is the heat flow in the x direction.
When dx→0, Eq. (I.7) becomes:
dqx φ
Lhx ⋅ δ fin ⋅ ⋅ dx = − 1 ⋅ dx . (I.8)
dx l fin
Rewrite Eq.(I.8) as:
dqx φ1
=− . (I.9)
dx l fin Lhx ⋅ δ fin
By Fouriers law, the conduction within the fin is expressed as:
dTx ( x)
q x = −Κ s , (I.10)
dx
where Кs is the conductivity of the solid material of the fin and Tx(x) is the
temperature distribution along x.
Substitution of Eq. (I.10) into Eq. (I.9) yields:
d 2Tx ( x) φ1
= . (I.11)
dx 2
l finδ fin Lhx Κ s
Therefore, it can be written as:
dTx ( x) φ1 ⋅ x
= + C1 . (I.12)
dx l finδ fin Lhx Κ s
By using the boundary condition:
dTx ( x)
at x = l fin , =0 (I.13)
dx
the coefficient in Eq. (I.12) is obtained as:
φ1
C1 = − . (I.14)
δ fin Lhx Κ s
By using Eq.(I.14) and (I.12), the temperature distribution can be written as:
φ1 ⋅ x 2 φ1 ⋅ x
Tx ( x) = − + C2 . (I.15)
2l finδ fin Lhx Κ s δ fin Lhx Κ s
By using the second boundary condition:
at x = 0 , Tx = Troot , (I.16)
the coefficient in Eq. (I.15) is obtained as:
Appendix 245

C2 = Troot . (I.17)
The temperature distribution is given as:
φ1 ⋅ x 2 φ1 ⋅ x
Tx ( x) = − + Troot . (I.18)
2l finδ fin Lhx Κ s δ fin Lhx Κ s
The temperature at the top of the fin is expressed as:
φ1 ⋅ l 2fin φ ⋅l
Tx x = l = − 1 fin + Troot . (I.19)
fin
2l finδ fin Lhx Κ s δ fin Lhx Κ s
The temperature difference between the top and the root of the fin is given as:
φ1 ⋅ l fin
∆T fin = Tx x = l − Troot = . (I.20)
fin
2δ fin Lhx Κ s
Substitution of Eq. (I.4) and (I.5) into (I.20) yields:
2Φ l 2fin  2 y0 
∆T fin = 2 ⋅  + 1 . (I.21)

πd hxΚ s Lhx  δ fin 

By numerical simulation for design, the heat exchanged at the ambient heat
exchanger is around 243 W, i.e. Φ=243. The designed diameter of the heat
exchanger is 29.7 mm. The solid material is choosen as copper, whose conductivity
is taken as 400 W/(m·K). Substitution of these properties into Eq. (I.21) yields:
l 2fin
∆T fin = 1116 ⋅ . (I.22)
Lhx
The length of fins is 4 mm, i.e. lfin=4 mm. Therefore, by Eq. (I.22), the temperature
span across the fin would be in the range of 12.5 to 33 K, if the length of the heat
exchanger is in the range of 0.534 mm to 1.425 mm as calculated at the beginning.
If so, the temperature span would be relatively large for a fin. A heat exchanger
with length of 0.534 mm to 1.425 mm would be a hard task for manufacturing, and
be fragile in use. By all these considerations, the length of the ambient heat
exchanger was designed to be 6 mm, which makes the temperature span across the
fin as small as 2.9 K.
246 Appendix

J: Time evolution of two orientations: upward


and downward in traveling-wave refrigerator
measurement

Here, the time evolutions for two orientations of the loop section—upward and
downward, are given to show that the loop downward configuration is more stable.
The temperature distribution of T1 to T11 is given in Fig. 4.5.6. During the
operations, the power into the PWG (P_driver in the Figs.J.1 and J.2) was
increased step-wise. For every input power to the PWG, the temperatures from T1
to T11 are traced and recorded. For Figs. J.1 and J.2, the left vertical axis is for
temperature (°C) and the right vertical axis is for power (W).
In Figs J.1 and J.2, the temperature T1, T2, and T5-T11 are stable, or “flat” for a
constant P_driver. Temperatures T3 and T4 are measurement of gas temperature
through and the solid temperature on the cold end heat exchanger. In Fig.J.1, T3
and T4 fluctuate slightly when P_driver is 30W. When P_driver is 40 and 50 W,
T3 and T4 show a difference between each other. When P_driver is more than 60
W, T3 and T4 start to fluctuate much.

35 100

30 90
25
80 T1
20 T2
70 T3
15
T4
Temperature (C)

10 60
T5
Power (W)

T6
5 50
T7
0 40 T8
0 500 1000 1500 2000 T9
-5
30 T10
-10 T11
20 P_driver
-15

-20 10

-25 0
Time (s)

Figure J.1: Time evolution of the loop upward configuration.


Appendix 247

35 100

30 90
25 T1
80
T2
20
70 T3
15 T4
Temperature (C) 60 T5
10

Power (W)
T6
5 50
T7
0 40 T8
0 400 800 1200 1600 2000 2400 2800 3200 T9
-5
30
T10
-10
20 T11
-15 P_driver
-20 10

-25 0
Time (s)

Figure J.2: Time evolution of the loop downward configuration.

In Fig.J.2, T3 and T4 show stable values, and the values of T3 are consistent with
those of T4 when P_driver is 30, 40 and 50 W. When P_driver is more than 60 W,
T3 and T4 start to fluctuate much.
By comparison of Figs. J.1 and J.2, it can be concluded that the loop downward
configuration is more stable than the loop upward. Therefore, in the measurements
of section 4.5, the loop downward configuration was employed.
248 Appendix

K: Acoustic field in the scaled-down standing-


wave systems
The resonator tube of a standing-wave system containing a parallel-plate stack is
schematically given in Fig. K.1. The resonator tube is divided into three zones by
the stack:
Zone 1: from the interface between the driver and the resonator tube to the left end
of the stack, it has a length of L1. The acoustic pressure and volume velocity at the
interface with the driver are p1in and U1in.
Zone 2: stack is placed between xL and xR, and having a length of Lstack. The stack
is sandwiched by cold and hot heat exchangers. The effects of two heat exchangers
on the acoustic field are neglected, assuming that they are ideal.
Zone 3: between the end of the resonator tube and the right end of the stack, it has
a length of L3. The acoustic pressure and volume velocity confined by the
boundary condition of the end of the resonator tube are p1end and U1end. The position
where the boundary condition applies is xend.

Lstack L3
L1
p1in p1end
U1in U1end

o x xL xR xend

Figure K.1: A sketch of the resonator tube of a standing-wave system.

Boundary conditions at xend:


Closed end: if the resonator tube is closed at xend, the solid wall imposes the
volume velocity at this position zero, i.e. U1end=0. Then the acoustic impedance at
this position is given by
p1end
Z end = = ∞. (K.1)
U1end
Open end: if the resonator tube is connected to a big gas reservoir, the open end
imposes the acoustic pressure at this position zero, i.e. p1end=0. Then the acoustic
impedance at this position is given by
p1end
Z end = = 0. (K.2)
U1end
Appendix 249

In zone 1 and 3, the tube segments are assumed as wide-open channel tube, i.e. the
hydraulic radius rh >> δ v and δ κ , thus Rott’s functions f v ≈ 0 and fκ ≈ 0 . Using
Eq. (G.10) in appendix G, substitution of Zc=Zend and xc-xb=L3 into Eq. (G.10)
yields:
ρma
Z end cos[kL3 ] + i sin[kL3 ]
ρ ma Ares
ZR = . (K.3)
Ares ρ m a cos[kL ] + iZ sin[kL ]
3 end 3
Ares
Ares is the cross sectional area of the resonator tube. Wave number k is defined as
Eq. (E.8) in appendix E:
ω
k = kr = . (K.4)
a
Substitution of Eq. (K.1) into (K.3) gives the acoustic impedance at xR when xend is
a closed end:
ρ m a cos[kL3 ]
Z R = −i . (K.5)
Ares sin[kL3 ]
Substitution of Eq. (K.2) into (K.3) gives the acoustic impedance at xR when xend is
an open end:
ρ m a sin[kL3 ]
ZR = i . (K.6)
Ares cos[kL3 ]
Similarly, the acoustic impedance at a random position x3 between xend and xR in
zone 3 is given by:
ρ ma
Z end cos[k ( x end − x3 )] + i sin[k ( x end − x3 )]
ρma Ares
Z3 = . (K.7)
Ares ρ m a cos[k ( x − x )] + iZ sin[ k ( x − x )]
end 3 end end 3
Ares
In the zone 2, where the stack places, assumption of wide-open channel is not valid
anymore. The hydraulic radius of the stack channel has the same order of gas
penetration depth, i.e. rh ~ 2δ κ and 2δν , thus Rott’s functions f v and fκ are
computed and taken into account.
In the zone 2, the transmission equation (G.7) in appendix G is employed.
Substitution of Zc=ZR and xc-xb=Lstack into Eq. (G.7) gives the acoustic impedance
at xL:
ZR
cos[kLstack ] + i sin[kLstack ]
ZL Z con
= . (K.8)
Z con ZR
cos[kLstack ] + i sin[kLstack ]
Z con
250 Appendix

The constant impedance Z con is defined as Eq.(G.5) in appendix G:


ωρ m
Z con = .
(1 − fν )Agask (K.9)

Here Agas is the cross-sectional area of the gas in the stack region.
The complex wave number in stack zone k is given in appendix E as:
ω 2 1 + (γ − 1) f k
k =
2
, (E.5)
a2 1 − fv
ω 1 + (γ − 1) f k
and k 1, 2 = ± . (E.6)
a 1 − fv
Similarly, the acoustic impedance at a random position x2 between xL and xR in
zone 2 is given by:
ZR
cos[k ( x R − x2 )] + i sin[k ( x R − x2 )]
Z2 Z con
= . (K.10)
Z con ZR
cos[k ( x R − x2 )] + i sin[k ( x R − x2 )]
Z con
Using transmission equation (G.10) in appendix G, and substitution of Zc=ZL and
xc-xb=L1 into Eq. (G.10) gives the acoustic impedance at x=0:
ρma
Z L cos[kL1 ] + i sin[kL1 ]
ρma Ares
Z0 = . (K.11)
Ares ρ m a cos[kL ] + iZ sin[kL ]
1 L 1
Ares
Similarly, the acoustic impedance at a random position x1 between xL and 0 in zone
1 is given by:
ρma
Z L cos[k ( x L − x1 )] + i sin[k ( x L − x1 )]
ρma Ares
Z1 = . (K.12)
Ares ρ m a cos[k ( x − x )] + iZ sin[k ( x − x )]
L 1 L L 1
Ares
In the scaled-down system, as shown in section 5.1, we have:
L1
Tube segments length: L1′ =
ϕx , (K.13)

′ = Lstack
Lstack
ϕx , (K.14)
L3
and L3′ =
ϕx . (K.15)

Similarly, for the random positions in three zones in scaled-down system become:
x′L − x1′ = ϕ x−1 (xL − x1 ) , (K.13a)
Appendix 251

x′R − x2′ = ϕ x−1 ( xR − x2 ) , (K.14a)


′ − x3′ = ϕ x−1 ( xend − x3 ) .
and xend (K.15a)
With Eq.(5.1.6), the wave number k ′ of the scaled-down system becomes:
ω′ ω
k′ = = ϕx = ϕxk . (K.16)
a′ a
The cross-sectional area of the resonator tube in scaled-down system becomes:
′ = (ϕ yϕ z )−1 Ares .
Ares (K.17)
The assumption that the ratio of stack spacing to thermal penetration depth is kept
constant while scaling, stated in Eq. (5.1.11a), holds here. So is the assumption that
the ratio between stack plate thickness and solid thermal penetration depth is fixed
while scaling, i.e. Eq. (5.1.12a). Thus the equations (5.1.8) to (5.1.18) are valid
here. Therefore, the Rott’s functions f v′ and f κ′ for the stack zone in scaled-down
system are given as stated in section 5.1:
y′
tanh (1 + i ) 0 ′  tanh (1 + i ) 0 
y
 δν   δν 
fν′ = = = fν , (5.1.16)

(1 + i ) y 0 ′ (1 + i ) y 0
δν δν
y′
tanh (1 + i ) 0 ′  tanh (1 + i ) 0 
y
 δ κ   δ κ 
f κ′ = = = fκ . (5.1.17)

(1 + i ) y 0 ′ (1 + i ) y 0
δκ δκ
With Eq.(5.1.11ab) and (5.1.12ab), the porosity of the stack in scaled-down system
is given:
∏′ y0′ ∏ y0
′ =
ψ stack = = ψ stack . (K.18)
∏′( y0′ + l ′) ∏( y0 + l )
The gas area Agas in scaled-down system is given:
A′gas = ψ stack
′ Ares′ = (ϕ yϕ z )−1 Aresψ stack = (ϕ yϕ z )−1 ⋅ Agas . (K.19)
Substitution of equations (K.16), (5.1.16) and (5.1.17) into (E.6) yields the
complex wave number in the stack zone of the scaled-down system:
ω ′ 1 + (γ − 1) f k′
k 1′, 2 = ± = ϕ xk 1, 2 . (K.20)
a′ 1 − f v′
Therefore, the substitution of equations (K.13) to (K.17), (5.1.16) and (5.1.17) into
the equations for acoustic impedance at xR (K.5) or (K.6) yields:
ρ m′ a′ cos[k ′L3′ ] ρma cos[ϕ x k ⋅ L3 / ϕ x ]
Z′R = −i = −i = (ϕ ϕ )Z ,
′ sin[k ′L3′ ]
Ares (ϕ yϕ z ) Ares sin[ϕ x k ⋅ L3 / ϕ x ] y x R
−1

(K.21)
252 Appendix

Or
ρ m′ a′ sin[ k ′L3′ ] ρma sin[ϕ x k ⋅ L3 / ϕ x ]
Z′R = i =i = (ϕ ϕ )Z .
′ cos[k ′L3′ ]
Ares (ϕ yϕ z ) Ares cos[ϕ x k ⋅ L3 / ϕ x ] y x R
−1

(K.22)
In the stack zone, the substitution of equations (K.13) to (K.22) into the equations
for acoustic impedance (K.8) and (K.9) yields the acoustic impedance of scaled-
down system at xL:
ω ′ρ m′ ϕ xωρ m
Z′con = = = (ϕ yϕ z )Z con ,
(1 − fν′ )Agas
′ k′ (1 − fν )(ϕ yϕ z )−1 Agasϕ xk (K.23)

Z′R
cos[k ′Lstack
′ ] + i sin[k ′Lstack
′ ]
Z′L Z′con
=
Z′con cos[k ′L′ ] + i Z′R sin[k ′L′ ]
Z′con
stack stack

(ϕ yϕ z )Z R cos[ϕ kL / ϕ ] + i sin[ϕ kL / ϕ ]
=
(ϕ yϕ z )Z con x stack x x stack x
Z
= L .
(ϕϕ Z )
cos[ϕ x kLstack / ϕ x ] + i y z R sin[ϕ xkLstack / ϕ x ]
Z con
(K.24)

(ϕ yϕ z )Zcon
Rewrite Eq.(K.24) as:
Z′con
Z′L = Z L ⋅ = (ϕ yϕ z )Z L . (K.25)
Z con
Similarly, in zone 1, substitution of equations (K.12) to (K.24) into the equations
for acoustic impedance (K.10) yields the acoustic impedance of scaled-down
system at x=0:
ρ m′ a′
Z′L cos[k ′L1′] + i sin[k ′L1′]
ρ m′ a′ ′
Ares
Z′0 =
′ ρ m′ a′
Ares cos[k ′L1′] + iZ′L sin[k ′L1′]

Ares
ρ ma
(ϕ ϕ )Z cos[ϕ x kL1 / ϕ x ] + i sin[ϕ x kL1 / ϕ x ]
=
ρma
y z L
(ϕ yϕ z )−1 Ares
(ϕ yϕ z )−1 Ares ρma
cos[ϕ x kL1 / ϕ x ] + i (ϕ yϕ z )Z L sin[ϕ x kL1 / ϕ x ]
(ϕ yϕ z )−1 Ares
= (ϕ yϕ z )Z 0 . (K.26)
Similarly, the acoustic impedances at random positions x1, x2 and x3 in scaled-
down system become:
Appendix 253

ρ m′ a′
Z′L cos[k ′( x′ L − x1′ )] + i sin[k ′( x′ L − x1′ )]
ρ m′ a′ ′
= (ϕ yϕ z )Z1 ,
Ares
Z1′ = (K.27)
′ ρ m′ a′
Ares ′ ′ ′ ′ ′ ′ ′
cos[k ( x L − x1 )] + iZ L sin[ k ( x L − x1 )]

Ares
Z′R
cos[k ′( x′R − x2′ )] + i sin[k ′( x′R − x2′ )]
Z′con
Z′2 = Z′con = (ϕ yϕ z )Z 2 , (K.28)
Z′R
cos[k ′( x′R − x2′ )] + i sin[k ′( x′R − x2′ )]
Z′con
ρ ′ a′
Z′end cos[k ′( xend
′ − x3′ )] + i m sin[k ′( xend ′ − x3′ )]
ρ m′ a′ ′
= (ϕ yϕ z )Z 3 .
Ares
Z′3 =
′ ρ m′ a′
Ares ′ ′ ′ ′ ′ ′ ′
cos[k ( xend − x3 )] + iZ end sin[k ( xend − x3 )]

Ares
(K.29)
When the input acoustic pressure at the interface between the driver and the
resonator tube p1in is constant in scaling, i.e.
p1′in = p1in , (K.30)
the volume velocity at the interface U1in is obtained as:
p1′in
= (ϕ yϕ z ) ⋅ U1in .
p1in −1
U1′in = =
Z′0 (ϕ yϕ z )Z 0 (K.31)

Therefore, the substitution of Eq. (K.31) into Eq. (K.27) gives the acoustic pressure
at a random position x1′ in the scaled-down system:
p1′ x ′ = x ′ = U1′in ⋅ Z1′ = (ϕ yϕ z ) U1in ⋅ (ϕ yϕ z )Z1 = p1 x = x .
−1
1 1
(K.32)
Similarly, the substitution of Eq. (K.31) into Eq. (K.28) and (K.29) gives the
acoustic pressures at random positions x′2 and x3′ in the scaled-down system:
p1′ x ′ = x ′ = U1′in ⋅ Z′2 = (ϕ yϕ z ) U1in ⋅ (ϕ yϕ z )Z 2 = p1 x = x ,
−1
2 2
(K.33)
p1′ x ′ = x ′ = U1′in ⋅ Z′3 = (ϕ yϕ z ) U1in ⋅ (ϕ yϕ z )Z 3 = p1 x = x .
−1
31 3
(K.34)
Therefore, the acoustic pressure remains unchanged locally at the same relative
position, when p1in is constant in scaling, i.e. p1′in = p1in . In another word, the
statement p1′ = p1 is valid everywhere in the scaled-down system.
Nomenclature

Lower case
p Pressure [Pa]
x Position along sound propagation [m]
y Position perpendicular to sound propagation [m]
u x component of velocity [m/s]
v y component of velocity [m/s]
s entropy per unit mass [J/(kg·K)]
a Sound speed [m/s]
cp Isobaric heat capacity per unit mass [J/(kg·K)]

cs Specific heat of the stack material [J/(kg·K)]

cV Isochoric specific heat [J/(kg·K)]

y0 Plate half-gap [m]

l Plate half-thickness [m]


e Energy per unit volume [J/m³]
f frequency [Hz]

fv Viscous Rott function —

fκ Thermal Rott function —


h Enthalpy per unit mass [J/kg]
i Imaginary unit —
m Molecular weight [kg/mol]
r Radius of tube [m]
d Diameter of tube [m]
t time [s]
k Wave number [m¯¹]
f (τ , bµ ) Function given by Eq. (F.21) —
Nomenclature 255

g (τ , bµ ) Function given by Eq. (F.22) —

Upper case
A Area [m²]
B Function given by Eq. (3.3.9) —
Function given by Eq. (3.3.10) or complex
D —
coefficients given by Eq. (F.38a) to (F.38d)
Coefficients of the solution of second order
C —
differential equation
T Temperature [K]
v
V Velocity [m/s]

E& Total energy flow [J/s]

W& Acoustic power flow [J/s]

Q& Heat flow [J/s]

K Thermal conductivity [W/(K·m)]


L length [m]
U Volumetric velocity [m³/s]
Re Real part of a complex function or number —
Im Imaginary part of a complex function or number —
I current [A]
V Voltage or volume [V] or [m³]
Z Acoustic impedance [Pa·s/m³]
F Force exerted on piston of driver [N]
Low-Reynolds-number-limit flow resistance of
R0 [Pa·s/m³]
regenerator at ambient temperature
C0 Basic compliance of regenerator [m³/Pa]
256 Nomenclature

Lower case Greek


β Thermal expansion coefficient [K¯¹]
γ Ratio, isobaric to isochoric specific heats —
ρ density [kg/m³]
µ Dynamic viscosity [Pa·s]
ξ the second viscosity [Pa·s]
ν Kinematic viscosity [m²/s]
εs Stack heat capacity ratio —

δκ Fluid’s thermal penetration depth [m]

δv Fluid’s viscous penetration depth [m]

δs Solid’s thermal penetration depth [m]

δ Wire diameter of stainless steel wire screen [m]


λ Wave length [m]
D Radian wavelength [m]
η efficiency —
κ Thermal diffusity [m²/s]
φ Phase angle —
ϕ Scaling factor —
σ Prandtl number —
ω Angular frequency [rad/s]
α Root of an equation —
θ Real coefficients given by Eq. (F.37a-d) —
τ Temperature ratio TH / TC —
ψ Porosity —

Upper case Greek


Γ Normalized temperature gradient —
Π perimeter [m]
Nomenclature 257

Σ Viscous stress tensor [N/m²]


ℜ Specific gas constant [J/(kg·K)]
ℜuniv Universal gas constant [J/(mol·K)]
Real coefficients given by Eq. (4.3.53), (4.3.54),
Θ —
(4.3.45) and (4.3.58)

Sub- and superscripts


crit Critical —
m Mean —
s Solid or standing —
κ thermal —
v viscous —
A amplitude —
h hydraulic —
H hot —
C cold —
1 First-order —
2 Second-order —

Others
~ Take conjugation —
‹ › Take space averaging —
¯¯¯ Take time averaging —
′ Scaled-down coordinate or y direction of solid —
→ vectors —
Bibliography

[1] N. Rott, “Thermoacoustics”, Advances in Applied Mechanics 20, 135


(1980).
[2] B. Higgins, Nicholson’s Journal I, 130 (1802).
[3] A.T. Jones, “Singing flames”, Journal of Acoustical Society of America 16,
254 (1945).
[4] A.A. Putnam and W.R. Dennis, “Survey of organ-pipe oscillations in
combustion systems”, Journal of Acoustical Society of America 28, 246
(1956).
[5] P.L.Rjike, “Notiz uber eine neue Art, die in einer an beiden Enden offenen
Rohre enthaltene Lift in Schwingungen zu verstezen”, Annln phys. 107,
339 (1859).
[6] K.T. Feldman, Jr., “Review of the literature on Rijke thermoacoustic
phenomena”, J. Sound Vib. 7, 83 (1968).
[7] C. Sondhauss, “Ueber die Schallschwingungen der Luft in erhitzten Glas-
Röhren und in gedeckten Pfeifen von ungleicher Weite”, Ann. Phys. 79, 1
(1850).
[8] C.T. Knipp, “A possible standard of sound”, Phys. Rev. 12, 491 (1917).
[9] K.W. Taconis and J.J.M. Beenakker, “Measurements concerning the vapor-
liquid equilibrium of solutions of He³ in He4 below 2.19K”, Physica 15, 733
(1949).
[10] T. Yazaki, A. Tominaga, and Y. Narahara, “Stability limit for thermally
driven acoustic oscillations”, Cryogenics 19, 490 (1979).
[11] Lord Rayleigh, “The explanation of certain acoustical phenomena”, Nature,
Lond. 18, 319 (1878).
[12] R.L. Carter, M. White, and A.M. Steele, Private communication of atomics
international division of north American aviation, Inc., September 24
(1962).
[13] K.T. Feldman, Jr., “A study of heat generated pressure oscillations in a
closed end pipe”, Ph.D. Dissertation, Mechanical Engineering, University
of Missouri (1966)
[14] K.T. Feldman, Jr., “Review of the literature on Sondhauss thermoacoustic
phenomena”, J. Sound Vib. 7, 71 (1968).
[15] W.E. Gifford and R.C. Longsworth, “Surface heat pumping”, Advances in
Cryogenic Engineering 1, 302 (1966).
Bibliography 259

[16] R. Radebaugh, “A review of pulse tube refrigeration”, Advances in


Cryogenic Engineering 35, (1990).
[17] R. Radebaugh, “Development of the pulse tube refrigerator as an efficient
and reliable cryocooler”, Proceedings of the Institute of Refrigeration 96, 1
(1999).
[18] I.A. Lyulina, R.M.M. Mattheij, A.S. Tijsseling, and A.T.A.M. de Waele,
“Numerical simulation of pulse-tube refrigerators”, International Journal of
Nonlinear Sciences and Numerical Simulation 74, 91 (2003).
[19] I.A. Lyulina, R.M.M. Mattheij, A.S. Tijsseling, and A.T.A.M. de Waele,
“Numerical simulation of pulse-tube refrigerators: 1d model”, Proceedings
of the Eurotherm Seminar: Heat Transfer in Unsteady and Transitional
Flows 5 (2004), 79-88, 287.
[20] P.Merkli and H. Thomann, “Thermoacoustic effects in a resonant tube”,
Journal of Fluid Mechanics 70, 161 (1975).
[21] H.A. Kramers, “Vibrations of a gas column”, Physica,’s Grav. 15, 971
(1949).
[22] G. Kirchhoff, “Uber den Einfluss der Warmeleitung in einem Gase auf die
Schallbewegung”, Annln. Phys. 134, 177 (1868).
[23] L. Trilling, “On thermally induced sound fields”, Journal of Acoustical
Society of America 27, 425 (1955).
[24] B.T. Chu, “Pressure waves generated by addition of heat in a gaseous
medium”, National Advisory Committee for Aeronautics, Tech. Note 3411
(1955).
[25] B.T. Chu, “Stability of systems containing a heat source-the Rayleigh
criterion”, National Advisory Committee for Aeronautics Res. Memo.
56D27 (1956).
[26] B.T. Chu and S.J. Ying, “Analysis of a self-sustained thermally driven
nonlinear vibration”, Physics Fluids 6, 1638 (1963).
[27] N. Rott, “Damped and thermally driven acoustic oscillations in wide and
narrow tubes”, Journal of Applied Mathematics and Physics 20, 230
(1969).
[28] N. Rott, “Thermally driven acoustic oscillations. Part II: Stability limit for
helium”, Journal of Applied Mathematics and Physics 24, 54 (1973).
[29] N. Rott, “The influence of heat conduction on acoustic streaming”, Journal
of Applied Mathematics and Physics 25, 417 (1974).
[30] N. Rott, “Thermally driven acoustic oscillations. Part III: Second-order heat
flux”, Journal of Applied Mathematics and Physics 26, 43 (1975).
[31] N. Rott and G. Zouzoulas, “Thermally driven acoustic oscillations. Part IV:
Tubes with variable cross-section”, Journal of Applied Mathematics and
260 Bibliography

Physics 27, 197 (1976).


[32] N. Rott and G. Zouzoulas, “Thermally driven acoustic oscillations. Part V:
Gas-liquid oscillations”, Journal of Applied Mathematics and Physics 27,
325 (1976).
[32] U.A. Müller and N. Rott, “Thermally driven acoustic oscillations. Part VI:
Excitation and power”, Journal of Applied Mathematics and Physics 34,
609 (1983).
[33] G.W. Swift, “Thermoacoustic engines”, Journal of Acoustical Society of
America 84, 1145 (1988).
[34] S. Backhaus and G.W. Swift, “A thermoacoustic-Stirling heat engine:
Detailed study”, Journal of Acoustical Society of America 107, 3148
(2000).
[35] S.L. Garrett, “Thermoacoustic engines and refrigerators”, American Journal
of Physics 72, 11 (2004).
[36] L.D. Landau and E.M. Lifshitz, “Fluid Mechanics”, Pergamon, Oxford,
1982.
[37] W.Y. Wu, “Fluid Mechanics” (in Chinese), Peking University press, 1982.
[38] G.W. Swift, “Thermoacoustics: A Unifying Perspective for Some Engines
and Refrigerators”, Acoustical Society of America, New York, 2002.
[39] L.E. Kinsler, A.R. Frey, A.B. Coppens, and J.V. Sanders, “Fundamentals of
Acoustics”, the fourth edition, John Wiley & Sons, Inc. 2000.
[40] G.W. Swift, “Thermoacoustic engines and refrigerators”, Physics Today,
22-28, (1995).
[41] J.C. Wheatley, T.J. Hofler, G.W. Swift, and A. Migliori, “An intrinsically
irreversible thermoacoustic heat engine”, Journal of Acoustical Society of
America 74, 153 (1983).
[42] J.C. Wheatley, T.J. Hofler, G.W. Swift, and A. Migliori, “Understanding
some simple phenomena in thermoacoustics with applications to acoustical
heat engines”, American Journal of Physics 53, 147 (1985).
[43] T.J. Hofler, “Thermoacoustic refrigerator design and performance”, Ph.D.
dissertation, Physics Department, University of California at San Diego,
(1986).
[44] S.L. Garrett, J.A. Adeff, and T.J. Hofler, “Thermoacoustic refrigerator for
space applications”, Journal of Thermophysics and Heat Transfer 7, 595
(1993).
[45] R.A. Johnson, S.L. Garrett, and R.M. Keolian, “Thermoacoustic cooling for
Surface Combatants”, Naval Engineers Journal 112, 335 (2000).
[46] J.C. Wheatley, G.W. Swift, and A. Migliori, “The natural heat engines”,
Los Alamos Science, 14, 2 (Fall 1986).
Bibliography 261

[47] J.A. Adeff and T.J. Hofler, “Design and construction of a solar,
thermoacoustically driven, thermoacoustic refrigerator”, Journal of
Acoustical Society of America (Acoustics Research Letters online)107,
L37 (2000).
[48] R.L. Chen, Y.C. Chen, C.L. Chen, C.L. Tsai, and J. DeNatale,
“Development of miniature thermoacoustic refrigerators”, 40th AIAA
Aerospace Sciences Meeting and Exhibit, January 14-17, 2002/Reno, NV.
[49] O.G. Symko et al. “High frequency thermoacoustic refrigerator”, US Patent
No. 10/458,752. (2004).
[50] O.G. Symko, “Miniature thermoacoustic refrigerator”, Final report Utah
University, Department of Physics, 1999.
[51] E.Abdel-Rahman, N.C. Azenui, I. Korovyanko, and O.G. Symko, “Size
considerations in interfacing thermoacoustic coolers with electronics”, 2002
inter Society Conference on Thermal Phenomena.
[52] E.M. Benavides, “Thermoacoustic nanotechnology: Derivation of a lower
limit to the minimum reachable size”, Journal of Applied Physics 101,
(2007).
[53] M.E.H. Tijani, “Loudspeaker-driven thermo-acoustic refrigeration”, Ph.D.
dissertation, Department of Applied Physics, Eindhoven University of
Technology, (2001).
[54] G.W. Swift, “LANL thermoacoustic animations”, Los Alamos National
Laboratory, http://www.lanl.gov/thermoacoustics/movies.html.
[55] I. Urieli and D.M. Berchowitz, “Stirling cycle engine analysis”, Hilger,
Bristol, UK, (1984).
[56] A.J. Organ, “Thermodynamics and gas dynamics of the Stirling cycle
machine”, Cambridge, (1992).
[57] C.D. West, “Liquid piston Stirling engines”, Van Nostrand Reinhold, New
York, (1983).
[58] P.H. Ceperley, “A pistonless Stirling engine—the traveling wave heat
engine”, Journal of Acoustical Society of America 66, 1508 (1979).
[59] P.H. Ceperley, “Gain and efficiency of a short traveling wave heat engine”,
Journal of Acoustical Society of America 77, 1239 (1985).
[60] T. Yazaki, A. Iwata, T. Maekawa, and A. Tominaga, “Traveling wave
thermoacoustic engine in a looped tube”, Phys. Rev. Lett. 81, 3128 (1998).
[61] S. Backhaus and G.W. Swift, “A thermoacoustic Stirling heat engine”,
Nature (London) 399, 335-338 (1999).
[62] A.M.Fusco and G.W.Swift, “Two-sensor power measurements in lossy
ducts”, Journal of Acoustical Society of America 91, (1992).
[63] H.Boogaart, Application of different regenerator materials and geometries
262 Bibliography

in a thermoacoustic engine. TUE report, March 2008.


[64] E.M.T.Moers, Regenerator materials for traveling wave thermoacoustic
systems. TUE report, September 2008
[65] A.A.Atchley, T.J.Hofler, M.L.Muzzerall, M.D.Kite and C.N.Ao,
“Acoustically generated temperature gradients in short plates”, Journal of
Acoustical Society of America 88, 251-263 (1990).
[66] Kim YT, Suh SJ, and Kim MG, “Linear resonant duct thermoacoustic
refrigerator having regenerator stacks”, In: Proceedings of the 16th
International Congress on Acoustics, June 1998, 821-822.
[67] A. Piccolo and G. Cannistraro, “Convective heat transport along a
thermoacoustic couple in the transient regime”, Int. J. Thermal Sci. 41,
1067-1075 (2002)
[68] A. Piccolo and G. Pistone, “Computation of the time-averaged temperature
fields and energy fluxes in a thermally isolated thermoacoustic stack at low
acoustic Mach numbers”, Int. J. Thermal Sci. 46, 235-244 (2006)
[69] D. Marx, Ph. Blanc-Benon, “Numerical calculation of the temperature
difference between the extremities of a thermoacoustic stack plate”,
Cryogenics 45, 163-172 (2004)
[70] A.T.A.M de Waele, “Basic treatment of onset conditions and transient
effects in thermoacoustic Stirling engines”, Journal of sound and vibration,
325(4-5), 974-988 (2009)
[71] W.C.Ward and G.W.Swift, “Design environment for low amplitude
thermoacoustic engines (DeltaE)”, J. Acoust. Soc. Am. 95, 3671-3672
(1994). Software and user’s manual also available from the Los Alamos
National Laboratory web site http://www.lanl.gov/thermoacosutics
[72] Web site http://sageofathens.com/
[73] W.Dai, E.C.Luo, Y.Zhang and H.Ling, “Detailed study of a traveling wave
thermoacoustic refrigerator driven by a traveling wae thermoacoustic
engine”, J. Acoust. Soc. Am. 119, 2686-2692 (2006).
[74] T.J.Hofler, “Improved efficiency and power density for thermoacoustic
coolers”, annual summary report, Naval postgraduate school, Monterey
California, June 1994.
[75] T.J.Hofler, “Improved efficiency and power density for thermoacoustic
coolers”, annual summary report, Naval postgraduate school, Monterey
California, July 1995.
[76] T.J.Hofler, “Improved efficiency and power density for thermoacoustic
coolers”, report, Naval postgraduate school, Monterey California, June
1996.
[77] R.S.Wakeland and R.M.Keolian, “Effectiveness of parallel-plate heat
Bibliography 263

exchangers in thermoacoustic devices”, J. Acoust. Soc. Am. 115 (6), 2873-


2886 (2004).
[78] M.E.H.Tijani and S.Spoelstra, “Study of a coaxial thermoacoustic-Stirling
cooler”, Cryogenics, 48, 77-82 (2008)
[79] E.C.Luo, W.Dai, Y.Zhang and H.Ling, “Thermoacoustically driven
refrigerator with double thermoacoustic-Stirling cycles”, Applied Physics
letters 88, 074102 (2006).
[80] E.C.Luo, G.Y.Yu, S.L.Zhu and W.Dai, “A high frequency
thermoacoustically driven thermoacoustic-Stirling cryocooler”,
Cryocoolers 14, International cryocooler conference, Inc. Boulder, 2007.
[81] Y.Ueda, T.Biwa, U.Mizutani, and T.Yazaki, “Experimental studies of a
thermoacoustic Stirling prime mover and its application to a cooler”, J.
Acoust. Soc. Am. 115 (3), 1134-1141 (2004).
[82] Y.Ueda, T.Biwa, U.Mizutani, and T.Yazaki, “Acoustic field in a
thermoacoustic Stirling engine having a looped tube and resonator”,
Applied Physics letters, volume 81, No.27, 5252-5254 (2002).
[83] Y.Ueda, T.Biwa, T.Yazaki, and U.Mizutani, “Construction of a
thermoacoustic Stirling cooler”, Physica B329-333, 1600-1601(2003).
[84] Paul Aben, “High-Amplitude thermoacoustic flow interacting with solid
boundaries”, Ph.D. dissertation, Department of Applied Physics, Eindhoven
University of Technology, (2010).
[85] I.Urieli and D.M.Berchowitz, “Stirling cycle engine analysis”, Hilger,
Bristol, UK (1984).
[86] J.R.Olson and G.W.Swift, “Similitude in thermoacoustics”, J. Acoust. Soc.
Am. 95 (3), 1405-1412 (1994).
[87] P.Lotton, P.Blanc-Benon, M.Bruneau, V.Gusev, S.Duffourd, M.Mironov
and G.Poignand, “Transient temperature profile inside thermoacoustic
refrigerators”, Int. J. Heat Mass Transf. 52 , 4986-4996 (2009).
[88] M.E.H.Tijani, S.Spoelstra, and P.W.Bach, “Thermal-relaxation dissipation
in thermoacoustic systems”, J. Applied Acoustics, 65, 1-13 (2004)
Summary
Thermoacoustics, as the word spelling indicates, is an interdisciplinary field in
physics. Both acoustics and thermodynamics are involved in the description of this
interesting phenomenon. When a solid wall is present in direction of the wave
vector of an acoustic field, the interaction between the wall and the acoustic wave
generates a transfer of heat from one location of the wall to another. This
mechanism can be used to drive so called standing- or traveling-wave type of
refrigerators and heat pumps. This is possible provided that instead of a solid wall a
heat transfer medium which essentially is a porous plug, is applied to increase the
effective surface area for the heat transport. Such a porous plug is known as stack
for standing wave type of apparatus, and regenerator in the case of traveling wave
type of apparatus. When the mechanism is used in the reverse way - an applied
temperature gradient along a wall - , an acoustic wave is generated spontaneousley
by the interaction between the sound wave and a solid. This is possible provided
that the temperature gradient exceeds a certain critical value, and by properly
designing a resonator and stack geometry to amplify one specific frequency. With a
mechanic-electric converter a so called, prime mover power generator can be made.
When a regenerator, which has a much smaller pore size than the stack, is
employed in an acoustic network, the Stirling cycle can be realized. The so-called
traveling-wave type refrigerators or engines are machines that use thermoacoustic
effects to realize the Stirling cycle.
The advantages of thermoacoustic machines over conventional refrigerators and
engines, such as no moving parts, environment-friendly, and high reliability, make
thermoacoustic apparatus of interest for future applications. Much work has been
done on larger-scale (1 - 25 meter) thermoacoustic machines in the past decades.
With the growing demand for small scale cooling devices in for instance laptop
computers, mobile phones, and satellites, development of miniature refrigerators
draws more and more attention. This PhD work is therefore dedicated to the
analysis of miniaturization of thermoacoustic refrigerators. Both types: standing-
wave and traveling-wave are investigated.
An introduction and historical review on thermoacoustics is given in chapter 1. In
chapter 2, the linear thermoacoustic theory is reviewed. The main content of
chapter 3 is about modeling of standing-wave systems and validation of the
modeling by investigation of a standing-wave type apparatus, which is similar to
the so-called “TAC” (thermoacoustic couple). At the end of chapter 3, the
measurements are analyzed and compared with computations based on proposed
modeling.
Summary 265

Chapter 4 is devoted to modeling traveling-wave systems and model validation by


experiments. An analytical model for a complete traveling-wave system of
refrigerator is developed. After that, the optimization of regenerator material is
considered. Various regenerator materials: metal honeycomb and ceramic
honeycomb with different channel shapes, stainless steel wire screens in different
hydraulic radii and porosities, were applied in a coaxial traveling-wave engine to
characterize their performance. The metal honeycombs gave partially because of
their larger cell size a poor performance, the ceramic honeycoms have an improved
performance, and the best performance was measured with the stainless steel wire
screens. The measurements of ceramic honeycombs and stainless steel wire screens
showed that there is an optimum value for the dimensionless hydraulic radius (the
ratio of hydraulic radius to a reference thermal penetration depth), which is around
0.3 for stainless steel wire screen regenerators and 0.16-0.2 for square cell
honeycombs. The efficiency goes up with increasing porosity, to a maximum value
above which the heat capacity of the solid becomes the limiting factor. The
conclusion for stainless steel wire screen regenerators is also in agreement with the
theoretical prediction by using the proposed analytical model. A full traveling-
wave refrigerator driven by a mechanical compressor was designed and built. This
system is again aiming at the validation of the model with measurements. The
theoretical computation showed an acceptable agreement with measurements.
In chapter 5, the analytical models for standing-wave systems (developed in
chapter 3) and traveling-wave systems (developed in chapter 4) are utilized to
characterize the performance in scaling down. By inserting three different scaling
factors: ϕ x , ϕ y , ϕ z , into the analytical models, the cooling power and efficiency
of scaling down systems are obtained. The results show that the scaling behaviour
of a standing-wave system is the same as that of a traveling-wave system. The
cooling power in the scaled-down system consists of two groups of energy flow
scaling with different factors: one group of energy flow scales with a
factor (ϕ yϕ z ) −1 , whereas the conduction loss scales as ϕ x (ϕ yϕ z ) −1 . Apparently, the
conduction loss term decreases less than the other one and results in a reduced
cooling power. So the efficiency decreases rapidly in scaling down. The term,
which is the product of the scaling factor and the original ratio of energy losses due
to thermal conduction to acoustic work, causes the reduction of efficiency after
scaling down. The unified scaling behaviour of standing-wave and traveling-wave
systems points out that the thermal conduction loss will finally dominate the losses
and becomes the limitation factor for scaling down.
We hope that the future design of mini-thermoacoustic-machine will benefit from
this finding to reduce the conduction loss, finally to develop smaller (< 5 cm) and
more powerful thermoacoustic refrigerators.
Samenvatting
Thermoakoestiek is zoals het woord al aangeeft een interdisciplinair veld van
onderzoek. Zowel akoestiek als thermodynamica spelen een rol in dit
gecompliceerde en interessante fenomeen. Als een wand aanwezig is in een
akoestisch geluidsveld treedt tussen de wand en het gas waar het geluidsveld heerst
een thermisch warmtepompproces op, waardoor warmte wordt verplaatst van een
plaats in de wand naar de andere. Met dit warmtetransportmechanisme kan een
staande of lopende golf type koelmachine of wamtepomp worden aangedreven. Dit
is mogelijk door gebruik te maken van een resonator. De randvoorwaarde daarbij is
dat voldoende wandoppervlak aanwezig is en dit wordt bereikt door de wand uit te
voeren als een poreuze plug. Deze poreuze plug wordt ook wel een stack of
regenerator genoemd, al naar gelang de poriegrootte en het type thermoakoestisch
apparaat. Als het mechanisme wordt omgekeerd - dus een temperatuurgradient
aangebracht langs een wand - kan een zwak akoestische golf in een resonator
versterkt worden, mits deze gradient een kritische waarde overschrijdt, dit wordt
een prime mover geneomd. Met een mechanisch-electrische omvormer kan de
akoestiek worden omgezet in electriciteit. Als een regenerator met een kleine
poriegrootte wordt ingezet als deel van een akoestische netwerk in een resonator is
het zelfs mogelijk om een Stirling cyclus te realizeren. Het is daarmee mogelijk om
een zogeheten "lopende golf" koelmachine te realiseren of een warmtemotor. Al
deze apparaten maken gebruik van een (benadering) van de Stirling cyclus.
De voordelen van thermoakoestische machines boven conventionele koelmachines
en motoren, zoals geen bewegende delen, milieuvriendelijk, en hoge
betrouwbaarheid, maken thermoakoestische apparaten van interesse voor
toekomstige toepassingen. In de afgelopen 20 jaar is veel onderzoek uitgevoerd aan
grootschalige thermoakoestische machines met lengteschalen van 2 tot 25 meter.
Met de groeiende vraag naar koelers in kleinschalige apparaten zoals satellieten,
mobiele telefoons, of laptop computers, komt er ook meer interesse voor de
ontwikkeling van miniatuur koelers waarbij thermoakoestiek een mogelijke
kandidaat kan zijn. Dit promotieonderzoek is daarom gericht op de analyse van
miniaturisatie van thermoakoestische koelapparaten. Zowel staande golf als
lopende golf types zijn onderzocht.
Een inleiding en historisch overzicht van thermoakoestiek wordt gegeven in
hoofdstuk 1. In hoofdstuk 2 wordt de lineaire theorie van thermoakoestiek
besproken. Dit vormt de noodzakelijke basis die nodig is om later in de
hoofdstukken 3, 4 en 5 de schaling te onderzoeken. Hoofdstuk 3 handelt over het
modelleren van staande golf systemen en het valideren van deze modellen met een
experimentele opstelling die bekend staat als een thermoakoestisch koppel. In het
Samenvatting 267

laatste gedeelte van hoofdstuk 3 worden de meetresultaten geanalyseerd en


vergeleken met modelberekeningen.
Hoofdstuk 4 gaat in op het modelleren van lopende golf systemen en model
validatie met experimenten. Een analytisch akoestisch model voor een complete
lopende golf machine is ontwikkeld. Voorts is de optimalisatie van regenerator
materiaal beschouwd. Verschillende honingraat regenerator materialen: metaal,
keramisch en met verschillende poriegrootte, alsook RVS metaalgaas met
verschillende hydraulische radius en porositeit zijn toegepast in een co-axiale
lopende golf machine om hun prestatie te karakteriseren. De metalen honingraten
gaven voornamelijk door hun te grote celgrootte de slechtste prestatie, de
keramische honingraten een gemiddelde prestatie en de beste prestatie wordt
geleverd door de metaalgaas regeneratoren. Uit de metingen is naar voren gekomen
dat voor zowel honingraat als gaasmateriaal er een optimale waarde is van de
hydraulische radius, welk 0.3 is voor gaasmateriaal en 0.16-0.2 voor honingraten
met vierkante celdoorsnede. Verder is gebleken dat de efficientie oploopt bij
toenemende porositeit tot een maximum dat wordt bereikt bij een porositeit van
ongeveer 90%, waarna de efficientie sterk afneemt vanwege de afnemende
warmtecapaciteit van het regeneratormateriaal in de limiet van een hoog poreuze
regenerator. Het analytische model laat een goede overeenkomst zien met de
resultaten van de metingen. Een lopende golf koelmachine die wordt aangedreven
door een mechanische compressor is ontworpen en gebouwd en metingen aan dit
systeem zijn uitgevoerd. Het is gebleken dat de metingen in redelijke
overeenstemming zijn met de modelvoorspellingen. De verschillen ontstaan door
thermische verliezen die niet zijn meegenomen in het model.
In hoofdstuk 5 zijn de analytische modellen voor staande en lopende golf systemen
gebruikt om de prestaties van thermoakoestische koelmachines te karakteriseren
voor miniaturisatie met schaalfactoren. Door een drietal schaalfactoren ϕx , ϕy , ϕz ,
(ϕ is de verhouding tussen het origineel en het miniatuurmodel) in de
thermoakoestische vergelijkingen in te voeren kan het gedrag van koelvermogen en
de efficientie van geschaalde systemen berekend worden. Het gedrag van lopende
en staande golf systemen is identiek. Het koelvermogen in het geschaalde model
bestaat uit twee groepen van energietermen met verschillende factoren.: een groep
die het akoestische vermogen levert schaalt met de factor , 1/(ϕy ,ϕz), terwijl de
andere groep van de warmtegeleiding schaalt met , ϕx/(ϕy ,ϕz). Bij een zuiver axiale
schaling nemen de verliezen van de warmtegeleiding naar verhouding dus toe, met
als gevolg een verminderd koelvermogen. De efficientie vermindert dus snel bij
miniaturiseren. Het universele schaalgedrag van thermoakoestische systemen toont
aan dat uiteindelijk de warmtegeleiding domineert en de limietfactor is bij
miniaturiseren. Toekomstige ontwerpen van thermoakoestische koelers op kleine
268 Samenvatting

schaal (< 5 cm) zijn daarom alleen mogelijk als de warmtegeleidingsverliezen


gereduceerd kunnen worden door andere regeneratormaterialen te kiezen.
Dankwoord
I would like to thank all people who participated in this project and my colleagues
whom I ever worked, and shared the enjoyable working atmosphere with.

I am very grateful to my daily advisor and co-promotor, Jos Zeegers, who coached
and supported me continuously during these five years of my PhD research. He
also helped me much in the starting phase of the PhD period to settle quickly in
Eindhoven. Furtermore I owe many thanks to my promotor prof. Marcel ter Brake
who initiated, and steered this project, and my second promotor prof. Fons de
Waele, both for all their expert guidance and support. Especially I like to mention
here the vast amount of time, and efforts they have spent in the phase of the
reading, correcting, and all support during the writing, of this thesis.
It should be mentioned here that this project was only possible through the kind
financial support of MicroNed, and this was made possible by Marcel's initiatives,
as well as the possibility to extend my contract with 8 months to finalize the
research work.
This work would not have been possible without the large technical support team,
who has realized all equipment. I am very thankful to our mechanical engineers
Leo van Hout, Paul Niël, and Henny Manders, who designed, and developed the
set-ups that were used in this research. Furthermore I have received much support
from our electronics engineers Peter Helfferich and Freek van Uittert, who
designed and built the electronics components often on day to day basis. Jørgen
van der Veen kindly helped with the photographs of parts of the set-up.
Much daily support in the actual manufacturing of the standing wave systems came
from the faculty workshop in cooperation with the support from team leader
Marius Bogers. I am thankfull to Henk van Helvoirt, Ginny Fransen, and Han den
Dekker who did the actual manufacturing.
Then the two large set-ups of the traveling wave systems have been built by the
University GTD workshops. Many technicians were involved in this. I like to thank
Mariëlle Dirks Smit who did the actual manufacturing via discharge machining of
the copper heat exchangers, Lucien Cleven and Jeroen Baijens turned, machined,
welded, and built the complex set-up. Jovita Moerel, and Jos van Kruijsdijk
designed, and built the electronical assembly of the set-up. Hans Wijtvliet together
with Freek van Uittert wrote all Labview codes to control the measurements. Erwin
Dekkers performed the strength computations, this all in support by team leader
Harrie de Laat.
My special thanks go to Paul Aben who shared the office with me for four years,
and helped me in translations of many letters I received in Dutch to English. I also
270 Dankwoord

want to thank my later office mate Christian Berendsen, who shared my room in
the last year in the office in Cascade. It was an enjoyable time for me to have lunch,
play poker games and chat with my former colleagues, Paul Aben, Wenqing Liang,
Paul Niël, in the old coffee room during lunch break.
For the stack materials we received much support from a number of people. I am
very thankful to prof. Chris Sutcliffe, Kaj Berggreen and Adam Clare, from the
Center of Materials of the School of Engineering of the University of Liverpool for
the supply of the stainless steel and plastic honeycombs. Furthermore I like to
thank John Wight of Corning USA, and Thierry Dannoux of Corning France for
the production of the large number of ceramic samples with high cell density,
which are in general not commercially available.
Then I like to thank Stan Lam from the Metaalgaasweverij Dinxperlo for his help
on supplying special wire gauze samples to be tested in the traveling wave set-up.
The work and design of the coaxial thermally driven traveling wave system was
possible thanks to the kind help of dr. Hassan Tijani, and I am very grateful for his
advice. Furthermore I like to thank Elise Moers who helped me with the
experiments on this engine.
In 2007 I could visit the Chinese Academy of Sciences in Beijing for a period of
three months. I am very thankful to dr. Wei Dai for his kind support, introducing
and helping me to learn the numerical thermoacoustic simulation tools developed
by him at CAS.
I would like to thank all my colleagues in Mesoscopic Transport Phenomena group,
whom I spent much joyful time with in playing bowling and dinners in prof. Mico
Hirschberg’s house. I would like to express my thanks to our secretary, Brigitte van
de Wijdeven, for taking care of all administrative details.
On a personal level, I want to show appreciation and gratitude to my husband and
my parents for their love and patience in these PhD years.
Curriculum Vitae
Yan Li was born in Liaoning, China, on February 14th 1977. After finishing her
pre-university education in hometown (Huludao in Liaoning province) in 1995, she
started her study, majoring in the speciality of aerospace engine in the department
of jet propulsion in Beijing University of Aeronautics and Astronautics (BUAA,
Beihang University, China) that same year. In the fourth year of her undergraduate
study, the year of 1999, she worked on project “Numerical Investigation of
Rotor/Stator Interaction Noise in an Aeroengine”. During the period of September
1999 to March 2002, she studied in field of aerospace propulsion theory and
engineering in the department of jet propulsion in BUAA and obtained her master’s
degree with thesis entitled “Sound Radiation Generated by Ducted Fan with
Supersonic Blade Tip Speed”.
From September 2003 till November 2005, she studied in McMaster University
(Canada) in the department of mechanical engineering and obtained master of
applied science with thesis entitled “Flow-Acoustics of T-Junctions: Effect of T-
Junction Geometry”.
From November 2005 till July 2010, she had been working as a PhD student at
Eindhoven University of Technology (the Netherlands) in the department of
applied physics. The project was mainly about scaling analysis of thermoacoustic
refrigerators, which was sponsored by MicroNed, and performed under the
supervision of prof.dr.ir. H.J.M.ter Brake, prof.dr.A.T.A.M.de Waele, and
dr.ir.J.C.H.Zeegers. She had the defense on her thesis, entitled “Thermoacoustic
Refrigerators: Experiments and Scaling analysis”, on the 27th of October 2011.

You might also like