Design of A Standing-Wave Thermoacoustic Engine: Catherine Gardner and Chris Lawn

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

DESIGN OF A STANDING-WAVE THERMOACOUSTIC

ENGINE
Catherine Gardner and Chris Lawn
School of Engineering and Materials Science Queen Mary, University of London, UK
email: c.j.lawn@qmul.ac.uk

The project entitled Stove for Cooking, Refrigeration and Electricity has the ambitious objec-
tive of making a 150 W electricity generator based on thermoacoustic principles that is driven
by the heat of a wood-burning stove. With a linear alternator involving the only moving part,
it is intended that the device should eventually be produced at very low cost for widespread
deployment in under-developed countries.
The first concept to be explored on the grounds of simplicity was the so-called standing-wave
design in which only a small fraction of the acoustic pressure is in-phase with the acoustic
velocity. The engine consisted of a closed linear duct containing a high-temperature heat
exchanger connected to the heat source and a low-temperature heat exchanger rejecting heat
to the surroundings, with a stack of narrow channels between them, and a linear alternator at
the low-temperature end of the duct. The high-temperature heat exchanger concept was one
of radiative transfer from a ceramic bulb in the fire to a number of metal plates spanning the
duct, and thence into the self-excited acoustic waves.
The conservation equations of thermoacoustics were solved together with equations describing
the radiative heat exchange, in order to determine the overall performance of this concept in
terms of the acoustic power that would be incident on a matched alternator at the chosen
frequency. Sensitivity studies were undertaken to examine the influence of the mean pressure
and of the composition of gas in the duct, and of the length and diameter of the passages in the
stack. Ultimately it was decided that the alternator could not be matched to the high acoustic
pressures involved (>100 mbar) and attention was switched to a travelling-wave device.

1. Introduction
A project entitled Stove for Cooking, Refrigeration and Electricity Generation is being un-
dertaken by a consortium of four UK universities and the charity Practical Action, with advice from
Los Alamos National Laboratory, and supported by the UK Engineering and Physical Sciences Re-
search Council under its International Development programme. The universities are Manchester,
Nottingham, City, and Queen Mary.
The project has the ambitious objective of making a robust electricity generator based on ther-
moacoustic principles, utilising the heat of a wood-burning stove. With a linear alternator involving
the only moving part, it is intended that the device should eventually be produced at very low cost
for widespread deployment in under-developed countries. At a later stage, the residual acoustic en-
ergy might be utilised to provide refrigeration capability in the same device. The project includes the

ICSV16, 59 July 2009, Krakow, Poland 1


16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

design of a stove to accommodate the thermoacoustic duct and to reduce the consumption of wood.
Stringent product design requirements have been developed to meet the perceived needs of a
particular village in Nepal [1], and these have been translated to design requirements for the engine.
Foremost among these in the current context are:

1. heat release from the combustion of the wood shall be at a rate of 4.4 kWth ;

2. electricity generation shall not interfere with the normal cooking process;

3. the thermoacoustic engine shall produce 150 We in full power operation;

4. the noise 1 m from the unit shall be <50 dBA;

5. the weight of the engine shall be <20 kg.

In addition, and of almost over-riding importance, the cost of the entire installation will be kept to a
level that is affordable by the target community.
This paper outlines the design of a device, details the method used to calculate its performance,
and reports some sensitivity calculations.

2. Design Concept
On the grounds of simplicity and robustness, the team first of all focussed upon the design of a
standing-wave engine. A possible realization of the concept, without the linear alternator, is shown
as Fig. 1. From the acoustic point of view, it comprised a closed-ended bounce volume, a hot-end
heat exchanger, a thermoacoustic stack, and a cold-end heat exchanger, terminating in a flange onto
which the linear alternator would have been bolted, all enclosed in a duct of approximately constant
cross-section. The bounce volume was necessary to allow useful magnitudes of acoustic velocity in
the stack, but it also provided the radiant surface for transmission to the hot-end heat exchanger. The
overall length of the duct was short compared with the acoustic wavelengths to be generated because
their frequency was determined by the resonant frequency of the alternator.

Linear Alternator
Radiant Bulb
Position

Stack

Hot Heat
Exchanger

Ambient Heat
Exchanger

Figure 1. Design concept for a standing-wave engine.

In order to avoid the complexity and expense of tube-sheets in the high temperature and possibly
sooty environment of the flue gases from the wood fire, it was decided to adopt a radiant heater design,

2
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

making use of the bounce volume. To give the required heat flux at the hot end of the thermoacoustic
stack with the envisaged maximum temperature for the radiant surfaces, it was necessary to have more
surface area than the cross-sectional area of the duct, so that there was adequate heat transfer on the
fire-side. This area was provided by pushing the radiant bulb horizontally through a port in the side
of the combustion chamber so that most of the surface of the bounce volume was transverse to the
flames: thus there were components of both radiant and convective heat transfer on that side. On
the duct side, the heat transfer was radiative, the heat exchanger plates seeing something close to
black-body radiation from the cavity, even if the emissivity of its surfaces was far from unity. The
heat exchanger plates were angled to prevent any direct view of the cavity by the stack, and had a
high absorptivity. They were set in the wall of the now-square duct and hence also conducted heat to
or from that wall.
The precise construction of the stack was not specified, but was likely to have been made of thin
metal sheets. The equivalent diameter and porosity were specified.
Although a tube-sheet design was clearly feasible for the cold-end heat-exchanger, and would
have been of similar construction to a car radiator, it was decided to investigate the capability of a
potentially cheaper and more robust approach in which heat transfer to a cold-water bath was purely
by conduction through a number of aluminium fins spanning the duct. Because the duct was to be
horizontal, it was convenient to mount it over a water-bath heat sink under the duct, and suffer the
penalty of essentially one-sided heat transfer through the fins.

3. The Thermoacoustic Model


The equations for conservation of mass, momentum and energy in 1-D form and the equation of
state were linearised with respect to acoustic perturbations, to generate model equations in the form
(see nomenclature, page 8):

v u dV
= jv = V u (1)
t x dx

u p
= ju = V {1 fv }
t x

tanh[(1+j) 4d ]
q
2
where fv (1+j) 4d
v
and v
(2)
v

V 1 T
Ta = p(1 fk ) fvk
Cp j x

h i
tanh (1+j) 4d
q
2 (1fk )Pr(1fv )
where fk = (1+j) 4d
k
and k
and fvk = (1fv )(1Pr)
(3)
k

p Ta v
= (4)
P T V
The unsteady momentum equation (2) includes the term fv for viscous dissipation, and that for energy
(3), the terms fk and fvk , in the forms expounded by Swift [2] for parallel plates. (Note that a different
definition of equivalent diameter from his is used here.) In the stack, it is deduced that the mean
temperature of the gas is locally also that of the stack surfaces, and that the enthalpy flux is invariant,
in the absence of significant 2-D effects. The mean pressure is also invariant, but all the gas properties

3
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

vary with the mean temperature, T . In the heat exchangers, a version of the standard fin conduction
equation with incident radiation

2T 00 A
kLt 2
= 2hhe (L + t)(Tg T ) + qrad , (5)
y Nw
allowing for convection to both sides and edges, can be solved to yield
cosh(my) q00 A
T Tg = (Tw Tg ) + rad (6)
cosh(mw/2) 2hhe N w(L + t)

where m2 = 2hkthe (1 + Lt ). The radiation flux is that normal to the duct cross-section and it has a view
factor VF for radiation onto the fins with absorptivity . For closed ends such as the bulb, VF will
only differ from unity if some of the radiation penetrates through the fins to the stack or the duct wall.
The radiation flux is given approximately by
00 4 4
qrad V F (Trad The ), (7)

since integration along the fin shows the average fin temperature to be
00 00
 
2 qrad A qrad A
The = Tg + Tw Tg tanh(mw/2) + (8)
mw 2hhe N w(L + t) 2hhe N w(L + t)
4
and average (The ) (average The )4 . No account is taken of the radiation penetrating the fins and
falling onto the end of the stack in the stack heat transfer analysis. The total heat transfer to the fins is
00
 
p q A
Q he = 2N 2hhe (L + t)kLt Tw Tg rad
tanh(mw/2) + Aqrad 00
(9)
2hhe N w(L + t)
The mean coefficients hhe for convective heat transfer from the fins are calculated from a correlation
taken from Piccolo and Pistone [3]. This is derived from the standard correlation for laminar flow
over a flat plate by integrating the instantaneous (spatially uniform) velocities over the cycle to yield
hhe Lef f
Nu = 0.507 Re0.5
L Pr
0.33
(10)
k
where the relevant velocity is the acoustic amplitude. Account must be taken of the fact that the
effective length for heat exchange may be less than the length of the fins in the flow direction so that
where is the local displacement amplitude. The numerical results generated by Piccolo
Lef f = 2,
and Pistone suggest that this correlation is pessimistic for d < 10 k , but there remains considerable
uncertainty in the correct hhe to use in the heat exchangers.

4. Solution Procedure
The total distance from the end of the radiant bulb to the flange connection for the linear alter-
nator is discretized into 1000 segments and the ends of the heat exchangers (and the stack) adjusted to
fall precisely on the nearest grid points. The calculation (in MATLAB) begins with a specification of
the acoustic pressure, impedance and frequency at the closed-end boundary, together with an estimate
of the mean gas temperature throughout the bounce volume, referenced to the fin root temperature.
The level of hot heat exchange is temporarily determined by a prescribed mean temperature gradient
at that end of the stack, specified as (T h T c )/L (stack length), where T c is a fictitious temperature,
referenced to T hec . T c is not the real cold-end gas temperature because the distribution through the
stack is not linear and is determined as the calculation proceeds. The thermo-acoustic equations are

4
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

solved sequentially on the grid. In the stack, iteration is necessary until a convergence criterion en-
suring H=0 is met at each point. The correlation of the current value of the acoustic temperature
fluctuation Ta , which depends on T x
, with the acoustic velocity u determines the enthalpy flux. Hence
we deduce,  
AC p uT a C p ufvk T
H = = Au (p(1 fk )) A u (11)
V V j x
so that a new temperature gradient can be determined from the enthalpy flux. This process is iterated
until the change in H is less than some chosen criterion. Typically, only two or three iterations are
required if there are several hundred mesh points in the stack. In the absence of a mean temperature
gradient (i.e for simple acoustic waves), no iteration is required to meet the criterion. In reality, there
may be a mean temperature gradient beyond the stack because of gas being displaced into the stack.
At the cold end, there is another change in enthalpy flux due to the change in mean temperature
gradient, the difference from the flux at the hot end being just the work done on the gas in adding
to the power of the acoustic wave. The calculation then proceeds to the end of the domain. These
calculations are repeated for a sequence of initial temperature gradients, i.e for a sequence of values
of T c , until the heat input from the cold end heat exchanger Q hec matches that calculated from the
change in enthalpy flux Q cold . This is achieved by putting
 !0.05
T h T c new Q cold
 = (12)
T h T c old Q hec
the exponent of 0.05 (or sometimes less) being chosen empirically. Having satisfied this boundary
condition, it is then necessary to match the hot-end heat exchange by iteration on the gas temperature
Th . This is done with another empirical relation
 !0.5
T h T heh new Q hot
 = (13)
T h T heh old Q heh

A new sequence of calculations with variable Tc is undertaken to match the cold end conditions again,
before adjusting the hot end Th once more, and so on until both conditions are satisfied. A complete
solution for one set of conditions has then been found, thus defining the impedance to which the
alternator would be matched. These calculations are then all repeated for a sequence of frequencies.

5. Results for a Typical Configuration


Results have been computed for a reference configuration with a cross-sectional area of 0.0225
2
m , a bulb length of 0.3 m, and heat exchangers with a depth of 0.1 m surrounding a stack of 0.3
m, followed by a short exit section. The porosity of the stack is 0.85 and the equivalent diameter of
the passages is 3 mm. Under the reference operational conditions, the frequency is 50 Hz, the mean
pressure, 10 bar, and the pressure amplitude at the end of the bulb, 0.5 bar. The axial distribution of
acoustic parameters is shown in Figure 2. The acoustic power transmitted from the open end of the
duct to an alternator presenting the correct impedance under these conditions is predicted to be 125
W, with a heat input of 1035 W, giving an efficiency of 12%.
Sensitivity calculations, in which just one parameter at a time was varied, were then undertaken,
with the results shown in Figures 3-6. The crucial parameters that explain most of these trends are
the ratios of penetration depth to equivalent passage diameter in the stack in equations (2) and (3).
Typically the penetration depths are < 0.3 mm at 50 Hz, but < 0.5 mm at 20 Hz, so less of the stack
passage is affected by the thermal wave as the frequency increases, and the acoustic power generation

5
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

Figure 2. Distribution of mean temperature, acoustic temperature, velocity and pressure.

Power added to the acoustic wave /W


150
Acoustic Power / W

100

50
20 25 30 35 40 45 50 55 60 65 70
Frequency / Hz
Power added to the duct /W
1200
Heating Power / W

1100

1000

900

800
20 25 30 35 40 45 50 55 60 65 70
Frequency / Hz

Figure 3. Variation of acoustic power and heating with frequency.

falls. In addition, viscous losses account for 50 W at 50 Hz, but only 5 W at 20 Hz. Thus, as frequency
increases, the efficiency of acoustic power generation falls. However, for the magnitude of acoustic
power, this is offset by a rise in the efficiency of heat exchange at the hot-end as the acoustic velocity
there increases (due to the shorter wavelength), negating the reduction in the displacement amplitude
u/ (Figure 3).
The effect of the increase in acoustic velocity that arises from the increase in specific volume
as the mean pressure is reduced for the same pressure amplitude (u pV /c) is similar (Figure
1/2
4), enhanced by P . However, there comes a point where the reduced density is unable
to sustain the required heat transfer without reducing the stack temperature gradient, and so there
is a broad optimum. Substituting helium for air in the duct gives the well-known enhancement in
performance due to the higher speed of sound and the greater acoustic velocity. Due to the increasing
dissipation of acoustic power in the long passages of the reference design, the increase in transmitted
power with pressure amplitude ( p2 for fixed temperatures in the stack) is reversed, the effect being
greatest for lower radiation temperatures and narrower passages (Figure 5). The optimum passage
width (d/2) depends heavily on the mean pressure, but is about 1 mm at 10 bar, and 2 mm at 3 bar,
or 4v in both cases (Figure 6). The optimum stack length for the reference operational conditions is
about 0.2 m, rather than 0.3 m.

6
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

700

600

500
Air, 923 K, 50 Hz
Helium, 923 K, 90 Hz
Acoustic 400
Power / W
300

200

100

0
1 2 3 4 5 6 7 8 9 10
P / bar

Figure 4. Variation of acoustic power with mean pressure.


250 250

200 200

150 150
Acoustic Acoustic
Power / W Power/ W
100 100

10 bar
50 Trad=1200 K, d=1.5 mm 50
Trad=1100 K, d=1.5 mm 3 bar
Trad=1200K, d=3 mm
Trad=1100K, d=3 mm
0 0
0 0.5 1 1.5 0 0.0025 0.005 0.0075 0.01
p^2 / bar^2 d/m

Figure 5. Variation of acoustic power with Figure 6. Variation of acoustic power with stack
pressure amplitude and radiation temperature. passage diameter and mean pressure.

6. Concluding Remarks
The exploration of the design space for this standing wave device, showed that there would
be a number of options for achieving the 150 We (>200 Wacoustic ) power target with otherwise the
reference geometry and conditions. This could be achieved with a pressure amplitude of 0.7 bar and
a stack equivalent diameter of 1.5 mm, for example. However, it was concluded that it would not be
possible to construct a linear alternator with the required characteristics: i.e. one with the capability of
withstanding the diaphragm rim stresses associated with the acoustic pressures. These are determined
by the absorbed power requirement, taken together with a stroke amplitude that would be limited to
20 mm. The acoustic pressures are so high for the required power because of the phase angle of about
88 associated with the standing wave design. The focus of the project has therefore now been
switched to a travelling wave device.

7. Acknowledgements
This work is supported by the EPSRC under Grant EP/E04462X/1. The contributions of the
partners in this project, particularly Scott Backhaus of LANL, are gratefully acknowledged.

7
16th International Congress on Sound and Vibration, 59 July 2009, Krakow, Poland

REFERENCES
1
P. H. Riley. and C. M. Johnson, Generating electricity in developing countries using thermo-
acoustics powered by burning wood, Acoustics 08, Paris, June 2008.
2
G. Swift, Thermoacoustic engines, J.Acoust. Soc. Am. 84(4),1145-1180, 1988.
3
A. Piccolo and G. Pistone, Estimation of heat transfer coefficients in oscillating flows: the ther-
moacoustic case, Int. J. Heat Mass Transfer, 49, 1631-1642, 2006.

Nomenclature T, Tw , Trad , Theh , Thec temperature of surfaces,


locally, at the duct wall, of the radiative
A cross-sectional area surface, and of width-wise averaged hot
Cp specific heat capacity and cold heat exchangers

c speed of sound T , T g , Th , T c mean temperature of the gas in the


stack, in the exchangers, and at hot and
d equivalent diameter, 4A/( perimeter) cold ends during interation

fv , fk , fvk thermoacoustic loss functions, equa- Ta acoustic temperature fluctuation


tions (2) and (3)
t time or thickness of exchanger plate
H enthalpy flux u acoustic velocity
hhe average mean heat transfer coefficient in V mean specific volume
heat exchanger
v acoustic perturbation of specific volume
k thermal conductivity
VF view factor of radiative surfaces from hot
L flow dimension of heat exchange plates heat exchanger entry plane
Lef f =L or 2 if 2 < L
w duct width
m parameter in fin conduction, equation (6)
x axial distance
N number of plates in heat exchanger
y transverse distance
Nu Nusselt number based on L
thermal diffusivity, kV /Cp
P mean pressure
v , k thermal penetration depths for momen-
p acoustic pressure tum, heat, eqs. (2) and (3)
Pr Prandtl number / emissivity of radiating surfaces

Q heh , Q hec total heat flux into the gas in hot, cold kinematic viscosity
heat exchangers, computed from hhe
acoustic displacement from mean

Qhot , Qcold total heat flux into the gas in hot, Stefan-Boltzman radiation constant
cold heat exchangers, computed from the
change in enthalpy flux circular frequency
00
qrad radiative heat flux normal to duct section overbar time mean

ReL Reynolds number based on Lef f and u amplitude

You might also like