Вы находитесь на странице: 1из 25

Journal of Sound and Vibration (1995) 180(4), 557581

THE CALCULATION OF THERMOACOUSTIC


OSCILLATIONS
A. P. D
Department of Engineering, University of Cambridge, Trumpington Street,
Cambridge CB2 1PZ, England
(Received 1 September 1992, and in final form 23 November 1993)
Thermoacoustic oscillations occur in a wide variety of practical applications in which
heat is supplied to an acoustic resonator. A simple geometry is investigated systematically
to determine the importance of various flow parameters on the frequency of the oscillations.
Detailed consideration of elementary examples shows that the form of the coupling between
the heat input and the unsteady flow has a crucial effect on the frequency of oscillation.
The same elementary examples are used to compare how well (if at all) different calculation
methods in the literature account for this influence. A mean flow and a distributed region
of heat input significantly complicate the analysis of thermoacoustic oscillations and are
often neglected. Model problems are used to illustrate that mean flow effects can become
significant even at modest inlet Mach numbers, and to indicate circumstances under which
a distributed heat input may be treated as concentrated.

1. INTRODUCTION

There is a possibility of thermoacoustic oscillations whenever combustion or heat exchange


take place within an acoustic resonator. They occur because unsteady heating generates
sound waves producing pressure and velocity fluctuations. However, within a resonator,
these in turn perturb the rate of heat input. Instability is possible if the phase relationship
is suitable because, while the acoustic waves perturb the heat input, the unsteady heat input
generates yet more sound.
Rayleigh [1] gave a clear physical description of this phenomenon. Acoustic waves gain
energy when the unsteady rate of heat input is in phase with the pressure perturbations.
If this energy gain exceeds that lost on reflection from the boundaries of the resonator,
linear acoustic waves grow in strength and the system is unstable. In practice, the
amplitudes of the pressure waves are limited by non-linear effects. Nevertheless, they can
become so intense that structural damage is done.
Many practical devices are susceptible to destructive thermoacoustic oscillations. They
occur in rockets, ramjets, gas boilers and aeroengines. Emphasis has been placed on
investigating the frequency of the oscillations and the mean heat input level for the onset
of the instability. Two excellent reviews of the literature have been given by Candel and
Poinsot [2] and Culick [3].
One particular thermoacoustic oscillation, involving the propagation of low frequency
longitudinal pressure waves, occurs in the afterburners of jet aeroengines. It is commonly
called reheat buzz. An experimental and theoretical investigation of this problem has
involved a series of tests on a laboratory-scale rig, designed to model the essential features
of an afterburner [4]. The rig geometry involves a flame burning in a duct in the wake of
a bluff body. There is a mean flow and the heat release is both distributed and unsteady.
It was found to be necessary to include a proper description of all these effects in the
557
0022460X/95/090557 + 25 $08.00/0

7 1995 Academic Press Limited

558

. .

development of a theory to explain the experimental results [5]. This is at variance with
views and alternative approaches reported in the literature, where one or more of these
effects are neglected.
The aim of this paper is to investigate systematically, for a very simple geometry, the
importance of various flow effects on the frequency of the thermoacoustic oscillations. This
is followed by a discussion on how well these effects are described by different calculation
methods.
Section 2 is concerned with the effect of the form of the coupling between the heat input
and the unsteady flow on the frequency of the combustion oscillation. In work on the
buzz oscillation [4, 5], this coupling was found to have a crucial effect on the frequency.
However, the complicated details of the experimental geometry and flow conditions
prevented any general conclusions from being drawn from this observation. A very simply
geometry is considered in section 2 to avoid this difficulty.
One-dimensional disturbances are investigated in a duct with no mean flow and one
closed and one open end. Unsteady heat is added at a single plane, across which the mean
temperature changes from T1 to T2 . Two particular forms for the relationship between the
rate of heat input and the flow are investigated in detail. Case I is that of no unsteady rate
of heat input. This represents a limiting case when the reaction time of the heating process
is much longer than the period of oscillation. Then, even though the fluid might enter the
heating zone unsteadily, the rate of heat input would be unable to adjust sufficiently rapidly
and would remain constant. In Case II the instantaneous rate of heat input is directly
proportional to the instantaneous mass flow rate into the heating zone: i.e., the heat input
per unit mass is constant. Such a heat input would be appropriate as a limiting case when
the reaction time of the heating process is so much faster than the period of oscillation that
the rate of heat input responds quasi-steadily to changes of flow rate into the heating zone.
Both forms of heat input are, of course, idealizations. However, it is worth noting that
the heat released by a non-premixed flame could approximate to Case I, when fuel is forced
into the combustion region at a constant rate and the combustion is directly proportional
to the fuel injection. Observations [5] of the rate of combustion for a premixed flame
burning in a duct are of the form of Case II at low frequencies, but with an additional
time delay.
The frequencies of the thermoacoustic oscillations are calculated in section 2.1 for these
two forms of the unsteady heat release rate. The relationship between the unsteady heat
input and the flow is found to crucially affect the frequency of oscillation. Indeed, by a
mean temperature ratio of six, there is nearly a 60% difference between the lowest
frequency predicted for Case II (when the instantaneous rate of heat input is proportional
to the instantaneous mass flow rate) and for Case I (when there is no unsteady rate of
heat input).
Some calculation methods commonly used in the literature take no account of the form
of the coupling between the unsteady heat input and the flow. In sections 2.2 and 2.3
respectively, it is shown that the calculation of the duct acoustic frequency and the Green
function technique recommended by Hedge et al. [6] are appropriate only when the rate
of heat input is not altered by the unsteady flow. The linearized Galerkin method [3]
acknowledges that the form of the coupling between the unsteady heat input and the flow
affects the frequency, but treats the shift in frequency, from that for no unsteady heat
release rate, as linear. This method is tested in section 2.4, by applying it to Case II, with
the instantaneous rate of heat input proportional to the instantaneous mass flow rate. The
predicted frequency and the true frequencies agree exactly for T2 = T1 , but diverge
significantly for larger temperature ratios. By a temperature ratio of six, the Galerkin
method predicts a frequency which is less that half the exact value.

559

The presence of a mean flow significantly complicates the analysis of a thermoacoustic


oscillation. An entropy wave, or convected hot spot, becomes coupled to the acoustic
waves and a simultaneous solution of equations of conservation of mass, momentum and
energy is required. Many authors choose to neglect the mean flow rather than deal with
this complication, their justification being that the inlet Mach number is low. The errors
introduced by such an approximation are investigated in section 3. It is striking how
significant Mach number effects can be. For example, one might be tempted to think that
an inlet Mach number of 015 is sufficiently small for the mean flow to be neglected. But
at this Mach number, the frequency of the thermoacoustic oscillation, for Case II,
T01 /T02 = 6, is reduced to half its no flow value.
Most practical ways of adding heat also exert a drag force on the flow. In work on the
buzz oscillation, the drag force exerted by the bluff-body flame holder was found not
to have a significant effect [7]. A systematic investigation into the effects of drag is given
in section 4. The drag force exerted by a grid or flame holder with a blockage ratio of 25%
or less is found to have a negligible effect on the frequency of thermoacoustic oscillations
 1 E 015. A blockage ratio of 50%
for the range of inlet Mach numbers considered, 0 E M
alters the frequency for inlet Mach numbers greater than 010.
When the heat release is distributed over an axial extent d rather than concentrated at
a single plane, the strengths of the acoustic and entropy waves vary continuously with
position. It is then easier to determine the unsteady flow directly by integrating the
equations of motion. That is done in section 5 to investigate the effects of an extended
region of heat input.
It is sometimes said that distributed heat input can be considered as concentrated
provided that its axial extent is very much smaller than the acoustic wavelength: i.e.,
provided that vd/c W 1, where v is the frequency and c is the speed of sound. However,
this is an oversimplification when there is a mean flow. Then there is also an entropy wave,
with the much shorter length scale u /v, where u is the mean flow speed. The entropy wave
is affected by even a modest spatial distribution of the heat input, and the frequency of
oscillation is altered significantly when this entropy wave is coupled to the acoustic waves.
For example, results in section 5 show that for an inlet Mach number of 010 and a
stagnation temperature ratio of nine, distributing the heat input over a length as short as
5% of the duct can lead to a 25% change in the lowest frequency of oscillation. To
guarantee, in general, that heat input over a region of length d can be treated as
concentrated requires not only vd/c W 1, but also vd/u W 1. The second inequality
is highly restrictive and circumstances under which it can be relaxed are discussed in
section 5.

2. THE EFFECT OF UNSTEADY HEAT INPUT

2.1.
In a region with heat input, the density r varies through changes in both the pressure
p and specific entropy s. The chain rule of differentiation shows that

Dr 1 Dp 1r Ds
=
+
.
Dt c 2 Dt 1s p Dt

(2.1)

When viscous and heat conduction effects are neglected, rT Ds/Dt = q(x, t), where q is
the heat input/unit volume and T is the absolute temperature. Moreover for a perfect gas,
1r/1s= p = r/cp = rT(g 1)/c 2, where c is the speed of sound, cp is the specific heat at

. .

560

constant pressure and g is the ratio of specific heats. Substitution into equation (2.1) leads
to

Dr 1 Dp
=
(g 1)q .
Dt c 2 Dt

(2.2)

Equation (2.2) may be applied to a combusting gas, provided that the reactants and
products behave as perfect gases, and there is no molecular weight change during the
chemical reaction [8].
For linear perturbations in a region where there is no mean heat input or mean velocity,
equation (2.2) can be linearized:

Dr 1 1p'
=
(g 1)q' .
Dt c 2 1t

(2.3)

The overbar denotes a mean value and the prime a perturbation.


A combination of equation (2.3) with the linearized equations of mass and momentum
conservation, Dr/Dt + r 9u' = 0 and 1u'/1t + 9p'/r = 0, leads to

0 1

1
(g 1) 1q'
1 1 2p'
.
r 9 9p' =
r
1t
c 2
c 2 1t 2

(2.4)

This inhomogeneous wave equation describes the pressure perturbation generated by the
unsteady heat input q'(x, t).
The solution of this wave equation can be readily calculated for one-dimensional
disturbances in the geometry illustrated in Figure 1. A duct of length l has one closed and
one open end. The heat input is harmonic of frequency v, concentrated at the plane x = b,
q'(x, t) = Q'(t)d(x b), where Q'(t) = Q
eivt and d denotes the Dirac d-function. In x Q b,
the mean temperature is T1 , the density r 1 and the speed of sound c1 . The solution of
equation (2.4), that satisfies the rigid end boundary condition at x = 0, is
p'(x, t) = A eivt(eivx/c1 + eivx/c1 ),
(2.5)
where the complex constant A has yet to be determined. Similarly, in x q b, the mean fluid
properties have values T2 , r 2 and c2 , and the solution of the wave equation can be written
in the form
p'(x, t) = B eivt(eiv(x l)/c2 eiv(x l)/c2 ),
(2.6)
after application of an open end boundary condition, p'(l, t) = 0. Integration of equation
(2.4), with q'(x, t) = Q'(t)d(x b), across the region x = b shows that
+

b
[p']xx =
= b = 0

and

$ %
1 1p'
r 1x

x = b+

x = b

(g 1) dQ'
,
r 1 c12 dt

since for a perfect gas, r 1 c12 = gp = r 2 c22. Equation (2.7b) is equivalent to


(g 1)
b+
[u']xx =
Q'(t),
= b =
r 1 c12

Figure 1. One-dimensional disturbances in a duct.

(2.7a, b)

(2.8)

561

a relationship between the volumetric expansion and the rate of heat input that has been
frequently used in the literature.
When Q'(t) is specified, the constants A and B may be calculated by substitution for
p'(x, t) from equations (2.5) and (2.6) into the jump conditions (2.7). The pressure
perturbation within the duct is then determined. However, thermoacoustic instabilities
occur when the heat input and the flow perturbations are coupled. As well as the unsteady
heating generating sound as described by equation (2.4), the flow perturbations in turn
affect the rate of heat input. The details of this feedback depend on the precise form of
the heating but, in all cases, lead to a relationship between Q
and the complex constants
A and B. This relationship, together with the jump conditions (2.7), constitute three
homogeneous equations for the three unknowns A, B and Q
, which can be written
symbolically in the form
A
X B = 0,
(2.9)
Q

23

where X is a 3 3 matrix. Such an equation will have a non-zero solution only if the
determinant of X vanishes, and this condition determines the frequency of the
thermoacoustic oscillation.
In this paper, two particular forms for the unsteady rate of heat input will be considered
as illustrative examples. The first case is that of no unsteady rate of heat input:
Case I,

Q'(t) = 0,

no unsteady rate of heat input.

(2.10)

Such a heat input would typically occur when the reaction time of the heating process is
much longer than the period of oscillation. Then, even though the fluid might enter the
heating zone unsteadily, the rate of heat input would be unable to adjust sufficiently rapidly
and would remain constant.
In the second case, the instantaneous rate of heat input is taken to be directly
proportional to the instantaneous mass flow rate into the heating zone, the heat input per
unit mass being that required to raise the mean temperature from T1 to T2 : i.e., cp (T2 T1 ).
Since the heat input per unit mass is constant, this case is referred to as that of no unsteady
heat input per unit mass. The rate of heat input is then proportional to the mass flow rate
into the heating zone:
Case II,

Q'(t) = cp (T2 T1 )r 1 u'1 , no unsteady heat input per unit mass,

(2.11)

where u'1 (t) denotes u'(b, t). Such a heat input would be appropriate when the reaction
time of the heating process is so much faster than the period of oscillation that the rate
of heat input responds quasi-steadily to changes of flow rate into the heating zone.
Of course, the two forms for the rate of heat input in (2.10) and (2.11) are idealizations.
However, it is worth noting that the heat released by a non-premixed flame could
approximate to Case I (expression (2.10)), if the fuel was forced into the combustion region
at a constant rate and the rate of combustion was directly proportional to the rate of fuel
injection. From a series of experiments on a confined premixed flame, Bloxsidge et al. [5]
postulated a rate of heat input which, for harmonic disturbances, can be expressed in the
form Q'(t) = kcp (T2 T1 )r 1 u'1 (t t); k and the time delay t depend on the frequency.
Over a range of measured frequencies, k varied between 05 and 10. There are, therefore,
also circumstances under which the Case II form of the heat input in expression (2.11) is
achievable in practice. The aim of this section is not to obtain detailed predictions for a
particular geometry of flame or heated grid, but rather to demonstrate that the form of
the unsteady heat input can have a significant effect on the frequency of a thermoacoustic

562

. .

oscillation. This leads on to an investigation of how this effect is accounted for (if at all)
in some of the methods of solution in the literature.
For the rate of heat input in Case I, described by equation (2.10), substitution for p'(x, t)
from equations (2.5) and (2.6) into the jump conditions (2.7) leads to an equation for v.
After some algebra the frequency of oscillation is found to be given implicitly by
tan (a) tan (b) = r 1 c1 /r 2 c2

for Case I,

(2.12)

where a = vb/c1 and b = v(l b)/c2 . It is a straightforward matter to solve equation (2.12)
numerically. The roots are real and the lowest frequency solution is shown in Figure 2 for
various values of T2 /T1 and for the particular geometry b = l/2.
When the rate of heat input in Case II described by equation (2.11) is used, a different
equation for the frequency is obtained. It is evident from equation (2.5) that
u'1 (t) = A eivt(eivb/c1 eivb/c1 )/r 1 c1 .

(2.13)

When this is used in equation (2.11), substitution into equation (2.7) leads to
tan (a) tan (b) = c1 /c2

for Case II.

(2.14)

Again the roots are real, and the lowest frequency solution is plotted in Figure 2 for
comparison with the solution of equation (2.12).
When the temperature is uniform along the duct (T2 = T1 ), the predicted frequencies just
correspond to that for which the duct length is a quarter-wavelength. For other
temperature ratios, the relationship between the unsteady heat input and the flow crucially
affects the frequency of oscillation. Indeed by a temperature ratio of six, there is nearly
a 60% difference between the frequency predicted for no unsteady heat input per unit mass
and no unsteady heat input per unit volume.
Of course, it is well known that the unsteady heat input affects the frequency. Rayleigh
[1] commented that the pitch is raised if heat be communicated to the air a quarter period
before the phase of greatest condensation: and the pitch is lowered if the heat be
communicated a quarter period after the phase of greatest condensation. In the geometry
considered here, the velocity fluctuation u'1 lags the pressure perturbation and so the rate
of heat input in equation (2.11) lags the greatest condensation and leads to a lower
frequency than that for Q' = 0. The numerical calculations show that this shift in frequency

Figure 2. The lowest frequency of oscillation as a function of temperature ratio T


 2 /T
 1 for b = l/2. , Case
I, no unsteady rate of heat input (roots of equation (2.12)); , Case II, no unsteady heat input per unit
mass (roots of equation (2.14)).

563

can be significant, and it is appropriate to discuss how this effect is accounted for in some
of the methods of solution in the literature.
2.2.
Many authors just assume that thermoacoustic oscillations occur at the acoustic
frequency of a duct. By this, they invariably mean the frequency at which a solution of
the homogeneous wave equation

0 1

1
1 1 2p'
r 9 9p' = 0
r
c 2 1t 2

(2.15)

satisfies the boundary conditions. This frequency takes no account of the relationship
between the unsteady heat release rate and the flow. For example, for the particular
geometry in Figure 1 it leads to
tan (a) tan (b) = r 1 c1 /r 2 c2 ,

(2.16)

for any form of the unsteady heat release rate. A comparison of equations (2.4) and (2.15)
shows why this is so. Equation (2.4), which was derived from the equations of motion,
reduces to the homogeneous form (2.15) only when there is no unsteady rate of heat input
per unit volume. When the rate of heat input is unsteady, the pressure perturbations satisfy
the inhomogeneous wave equation (2.4) and the form of the heat input must be considered
in a calculation of the frequency of oscillations in the duct.
2.3.
Hedge et al. [9] used a Green function technique to determine the one-dimensional
pressure perturbation generated by unsteady combustion within a duct. They essentially
solved the inhomogeneous wave equation (2.4) for q'(x, t) = q (x) eivt, by introducing a
Green function G(x=x0 ) which satisfies
r c 2

0 1

d 1 dG
+ v 2G = cp (g 1)d(x x0 )
dx r dx

(2.17)

and the boundary conditions. The solution of equation (2.4) is then given by
p'(x, t) = p (x) eivt, where
p (x) =

iv
cp

G(x=x0 )q (x0 ) dx0 .

(2.18)

In particular, when the heat input is concentrated on the plane x = b, q (x0 ) = Q


d(x0 b),
and
p (0) = ivQ
G(0=b)/cp .

(2.19)

A general form for the Green function was given by Hedge et al. [9, equation 14]. For
the geometry illustrated in Figure 1, their expression simplifies to
sin (b)
c1 c2
.
v T1 c2 sin (a) sin (b) T2 c1 cos (a) cos (b)

G(0=b) =

(2.20)

Hedge et al. found the resonance frequencies of the duct by minimizing the magnitude of
the denominator of the Green function. For the particular form (2.20), the denominator
vanishes at frequencies that satisfy
tan (a) tan (b) = T2 c1 /T1 c2 = r 1 c1 /r 2 c2 ,

(2.21)

564

. .

a condition identical to that in equation (2.12). If the unsteady heat input is a specified
broadband source, equation (2.19) shows that large pressure perturbations are generated
at this frequency. However, it is misleading to interpret this as the frequency of a
thermoacoustic instability, as suggested in reference [6].
A thermoacoustic instability involves coupling between the heat input and the flow. The
unsteady heat input occurs as a response to fluctuations in flow. This can be expressed
in the form
Q
= Z(v)p (0),
(2.22)
for some function Z(v). Substitution for Q
in equation (2.19) leads to
p (0)(1 ivZ(v)G(0=b)/cp ) = 0.

(2.23)

The frequency of the thermoacoustic oscillation is the frequency at which self-sustaining


oscillations can occur. It is clear from equation (2.23) that p (0) can be non-zero only if
{1/G(0=b)} ivZ(v)/cp = 0.

(2.24)

This reduces to 1/G(0=b) = 0 only if Z(v) = 0. This explains why the zeros of 1/G(0=b) were
found in equation (2.21) to be identical to frequencies calculated for the particular case
of no unsteady rate of heat input. When the heat input responds to the flow, Z(v) is
non-zero and the frequency is shifted.
As an example, consider Case II with no unsteady heat input per unit mass. The mode
shape in x Q b, as described by equations (2.5) and (2.13), shows that
u'1 (t) = i sin (a)p (0) eivt/r 1 c1 .

(2.25)

Hence the unsteady heat input described by equation (2.11) can be written in the form
Q
= icp (T2 T1 ) sin (a)p (0)/c1 , leading to
Z(v) = icp (T2 T1 ) sin (a)/c1 .

(2.26)

Substitution of the Green function and Z(v), from equations (2.20) and (2.26),
respectively, into equation (2.24) leads, after some algebra, to the same equation for v as
that given in equation (2.14).
In summary, a Green function can be a useful tool in the investigation of thermoacoustic
oscillations. However, it is not sufficient simply to inspect the Green function alone. The
form of the coupling between the heat input and the flow must also be considered. In the
notation of this section, this coupling is described by the function Z(v). The frequency
of the thermoacoustic oscillation is then a zero of {1/G(0=b)} ivZ(v)/cp . The form of
the relationship between the unsteady heat input and the flow affects the frequency.
2.4.
Culick and his colleagues used a Galerkin method to solve the inhomogeneous wave
equation (2.4). This method acknowledges that the form of the coupling between the
unsteady heat input and the flow affects the frequency, but linearizes the change in
frequency and mode shape. Reference [3] clearly describes the method and gives a review
of earlier work.
For the one-dimensional problem illustrated in Figure 1, the method involves expanding
the pressure perturbation as a Galerkin series:
a

p'(x, t) = s hm (t)cm (x).


m=1

(2.27)


The functions cm (x) are eigensolutions of the homogeneous wave equation,

0 1

vm2
d 1 dcm
c + r
= 0,
c 2 m
dx r dx

m = 1, 2, . . . ,

565

(2.28)

that satisfy the boundary conditions


dcm /dx = 0 at x = 0
and
cm = 0 at x = l.
(2.29)
It is a straightforward matter to show from this definition that the functions cm are
orthogonal,

cm (x)cn (x) dx = 0

for m $ n.

(2.30)

For the geometry in Figure 1, the mean temperature is uniform in the two regions x Q b
and x q b. The homogeneous wave equation in (2.28) then simplifies to
d2cm vm2
d2cm vm2
in 0 E x Q b,
+ 2 cm = 0 in b Q x E l, (2.31a, b)
2 cm = 0
2 +
dx
c1
dx 2
c2
together with two jump conditions across x = b:
+

b
[cm ]xx =
= b = 0

and

$ %
1 dcm
r dx

x = b+

= 0.

(2.32)

x = b

It is a matter of straightforward algebra to check that


cm (x) =

cos (vm x/c1 )


in x Q b
(cos am /sin bm ) sin (vm (l x)/c2 )
in x q b

(2.33)

in a solution of equations (2.31) and (2.32) that satisfies the boundary conditions (2.29).
vm is the mth root of tan am tan bm = (r 1 c1 )/(r 2 c2 ) and am = vm b/c1 and bm = vm (l b)/c2 .
Substitution for p'(x, t) from equation (2.27) into equation (2.4) leads to

s
m=1

d2hm
1q'
+ vm2 hm cm (x) = (g 1)
.
dt 2
1t

(2.34)

After multiplication by cn (x) and integration with respect to x, the orthogonality


condition (2.30) shows that equation (2.34) simplifies to
(g 1)
d2hn
+ vn2hn =
En
dt 2

1q'
c (x) dx,
1t n

(2.35)

where
En =

cn2 dx = 12 (b + (l b) cos2 am /sin2bm ),

(2.36)

after substitution for cn from equation (2.33).


No approximation has yet been made, but in the next stage in the analysis it is assumed
that 1q'/1t is small in magnitude. It is then argued that since 1q'/1t is small, it need only
be evaluated approximately. The acoustic approximations hn (t)cn (x) and (h n (t)/r vn2 ) dcn /
dx for the pressure and velocity perturbation are then used when calculating 1q'/1t. If
second derivatives of the amplitudes arise, they are replaced by the zeroth order
approximation h n (t) 2 vn2hn (t). The errors induced by these approximations can be
checked by applying the method to Case II to find the lowest frequency of combustion
oscillation when there is no unsteady heat input per unit mass.

. .

566

Then the rate of heat input has the form given by equation (2.11) and
q'(x, t) = cp (T2 T1 )r 1 u'1 d(x b). Rewriting the velocity perturbation in the acoustic
form (h 1 (t)/r v12) dc1 /dx leads to
q'(x, t) = cp (T2 T1 )

h 1 (t) dc1
(b ).
v12 dx

(2.37)

The right side of equation (2.35) therefore involves the second derivative, h 1 (t). This is
replaced by v12h1 (t), following Culicks rules, to give
h 1 + v12h1 = h1

cp
dc
(g 1)(T2 T1 ) 1 (b)c1 (b)
E1
dx

= h1 cp v1 (g 1)(T2 T1 ) sin a1 cos a1 /E1 c1 ,

(2.38)

after substitution for c1 from equation (2.33). Equation (2.38) is a straightforward second
order differential equation for h1 (t). The frequency of oscillation of h1 (t), v, can be found
by substituting h1 (t) = C eivt into equation (2.38) to give
v 2 = v12 cp v1 (g 1)(T2 T1 ) sin a1 cos a1 /E1 c1 .

(2.39)

As in this method it is assumed that the difference v v1 is small, Culick [3] therefore
recommended that the square root in equations such as (2.39) be evaluated approximately
by using the binomial theorem. Equation (2.39) then simplifies to
v = v1 cp (g 1)(T2 T1 ) sin a1 cos a1 /2E1 c1 .

(2.40)

This approximation to the lowest frequency of combustion oscillation is plotted in Figure


3 for comparison with the exact value.
The linearized Galerkin method calculates the change in frequency due to the unsteady
heat input, but treats the shift as small. The form of heat input in equation (2.11), i.e.,
no unsteady heat input per unit mass, vanishes identically for T2 = T1 and is infinitesimally
small for small T2 T1 . Hence, the frequency predicted from the linearized Galerkin
method and the true frequency agree exactly for T2 = T1 . The two curves of frequency
against temperature ratio also have the same slope at this point, but they diverge
significantly for larger temperature ratios, as shown in Figure 3. By a temperature ratio
of six, the Galerkin method predicts a frequency which is less than half the exact value.

Figure 3. The lowest frequency of oscillation for Case II with no unsteady heat input per unit mass for b = l/2.
, Correct solution; , acoustic resonance frequency and the value predicted by investigating zeros of
1/G(0=b); , frequency predicted by linearized Galerkin theory; , frequency predicted by ray theory.

567

2.5.
The phase change of a high frequency sound wave, as it travels through a region of
gradually varying sound speed c (x), can be described by ray theory. According to ray
theory, negligible sound energy is reflected as the ray propagates through such a region,
and a sound wave of frequency v, propagating in the x-direction undergoes a phase change
v f0l dx/c (x) over a distance l [10].
Ray theory is appropriate for short wavelength waves travelling through a region in
which the sound speed varies gradually and there is no unsteady heat input per unit
volume. It is therefore entirely the wrong physical limit for low frequency thermoacoustic
oscillations, where the wavelength is longer than, or comparable with, the length of the
region of heat input and the oscillations are driven by the unsteady heating. Nevertheless,
ray theory is used in industry to estimate the frequency of thermoacoustic oscillations and
so it is worthwhile to compare its predictions with the exact frequency in some model
calculations.
According to the ray theory, the phase of a ray is changed by v(b/c1 + (l b)/c2 ), after
travelling along the duct in Figure 1. The pressure wave undergoes phase changes at 0 and
p on reflection at the closed and open ends of the duct respectively. An estimate of the
resonance frequency then follows from a requirement that the total phase change after
propagation up and down the duct be an integer multiple of 2p: i.e.,
2v(b/c1 + (l b)/c2 ) + p = 2pn,

n integer.

(2.41)

This clearly gives frequencies which are independent of the form of the rate of heat release.
The lowest frequency solution of equation (2.41) is plotted in Figure 3. It agrees with
the true frequency of oscillation only for the trivial case with no mean temperature
gradient, T2 = T1 . When the mean temperature varies along the duct over a length scale
which is much shorter than the wavelength, ray theory is inadequate. This is because it
does not take reflections from the region of heat input into account and so is unable to
describe the flow perturbations correctly.
2.6.
Thermoacoustic oscillations have been investigated for the geometry illustrated in Figure
1. Results for the frequency of oscillation for two particular forms of the unsteady heat
release rate are summarized in Figure 2. They show that the form of the relationship
between the unsteady heat input and the flow crucially affects the frequency of oscillation.
Many authors have neglected this effect or treated it as small.
It has been said [3] that good approximations to the frequency of a thermoacoustic
oscillation can often be obtained by calculating the acoustic frequencies of the geometry.
That is equivalent to solving the case of no unsteady heat release rate per unit volume.
The results in Figure 3 serve as a warning that this procedure is not universally appropriate.
Hedge et al. [6] recommended the use of a Green function G to investigate
one-dimensional perturbations in a duct. They then interpreted the roots of 1/G = 0 as the
frequency of the thermoacoustic oscillation. In Figure 3 it is shown that this leads to results
which can be significantly in error for the particular case of no unsteady heat input per
unit mass. The analysis in section 2.3 explains why it is not sufficient simply to inspect
the Green function alone. The frequency of a thermoacoustic oscillation is a zero of 1/G
only if the rate of heat input per unit volume does not vary significantly in response to
the flow perturbations. If it does, the form of the coupling between the heat input and the
flow affects the frequency and must be considered.
Culick and his colleagues have developed a Galerkin method to investigate thermoacoustic oscillations. This method acknowledges that the form of the coupling between the

568

. .

unsteady heat input and the flow affects the frequency, but treats the shift in frequency,
from that for no unsteady heat release rate per unit volume, as linear. The method
therefore works exactly when there is no unsteady heat release rate per unit volume. In
section 2.4 the method was applied to the case of no unsteady heat input per unit mass
and once again the results are plotted in Figure 3. The predicted frequency and true
frequencies agree exactly for T2 = T1 . The two curves of frequency against temperature
ratio also have the same slope at this point, but they diverge significantly for larger
temperature ratios. By a temperature ratio of six, the Galerkin method predicts a frequency
which is less than half the exact value.
In section 2.5 the frequency predicted by ray theory was compared with the true
frequency. There is no theoretical basis for believing that ray theory can accurately describe
low frequency thermoacoustic oscillations. Indeed, the plots in Figure 3 show that ray
theory is unable to describe correctly the variation of the frequency of oscillation with the
temperature ratio T2 /T1 .

3. THE EFFECT OF A MEAN FLOW

Most practical thermoacoustic oscillations involve a mean flow. However, often the
Mach number of the oncoming flow is so small that it is tempting to neglect the mean
velocity entirely. The errors introduced by such an approximation are investigated in this
section.
A mean flow has two main consequences. Trivially, it affects the speed of the
propagation of the acoustic waves, which then travel downstream with speed c + u and
upstream at c u . In addition, a mean flow admits a second type of linear waves. These
are convected with the mean velocity and involve convected entropy or vorticity. As a
consequence, the equations of conservation of mass, momentum and energy become
coupled.
These effects may be illustrated by considering one-dimensional thermoacoustic
oscillations in the geometry in Figure 4. The inlet is choked so that the mass flow rate is
constant there:
(r'/r 1 ) + (u'/u1 ) = 0

at x = 0.

(3.1)

The boundary condition of a downstream open end can again [10] be taken as
p' = 0

at x = l.

(3.2)

The heat input is concentrated at the fixed plane x = b, the rate of heat input per unit
cross-sectional area being denoted by Q'(t).
Consider disturbances with time dependence eivt. Upstream of the zone of heat input,
there are acoustic waves propagating in both directions and the flow is isentropic. The

Figure 4. One-dimensional disturbances in a duct with a mean flow and mean rate of heat release.

569

perturbation in flow quantities can be written in the form


p'(x, t) = eivt(A eivx/c1 (1 + M 1 ) + B eivx/c1 (1 M 1 ) ),

(3.3a)

u'(x, t) = eivt (A eivx/c1 (1 + M 1 ) B eivx/c1 (1 M 1 ) )/r 1 c1 ,

(3.3b)

r'(x, t) = p'(x, t)/c ,


2
1

cp T'(x, t) = p'(x, t)/r 1 ,

(3.3c, d)

for 0 E x Q b, where A and B denote the strengths of the acoustic plane waves and
M
 1 = u1 /c1 .
Downstream of the region of heat input, there might be a convected hot spot in addition
to plane sound waves, and so
p'(x, t) = eivt (C eivx/c2 (1 + M 2 ) + D eivx/c2 (1 M 2 ) ),
ivt

u'(x, t) = e (C e

ivx/c2 (1 + M
2)

r'(x, t) =
cp T'(x, t) =

D e

ivx/c2 (1 M
2)

(3.4a)

)/r 2 c2 ,

(3.4b)

p'(x, t) Sr 2 iv(t x/u2 )

e
,
c22
cp

(3.4c)

p'(x, t)
Sc22
+
eiv(t x/u2 ),
r 2
(g 1)cp

(3.4d)

for b Q x E l. C and D are the amplitudes of the acoustic waves, S is that of the entropy
wave or convected hot spot and there are no vorticity waves in this one-dimensional
example. M
 2 = u2 /c2 .
The constants A, B, C, D and S are related by conservation conditions across the zone
of heat input. If the heat input is at a constant position x = b, then conservation of mass
across this discontinuity shows that
r 1 u'1 + r'1 u1 = r 2 u'2 + r'2 u2 ,

(3.5)

where the suffices 1 and 2 denote the flow quantities at x = b and b


Similarly, it follows from conservation of momentum across x = b that

respectively.

p'1 + r'1 u12 + 2r 1 u1 u'1 = p'2 + r'2 u22 + 2r 2 u2 u'2 .

(3.6)

The perturbation in the energy flux out of a control volume around x = b exceeds the
energy flux into it by Q'(t): i.e.,
cp T01 (r 1 u'1 + r'1 u1 ) + r 1 u1 (cp T'1 + u1 u'1 ) + Q'
=cp T02 (r 2 u'2 + r'2 u2 ) + r 2 u2 (cp T'2 + u2 u'2 ),

(3.7)

where T0 is the mean stagnation temperature T + u /cp .


Care needs to be taken to recover the jump conditions for zero mean flow from equations
(3.5)(3.7). The strength S of the entropy wave enters these equations only in the
combination u2 S. In the limit u2 :0, S tends to infinity, while the product u2 S remains
finite. It is convenient to begin an investigation of low mean flow Mach numbers by noting
that equation (3.5) may be used to recast the energy equation (3.7) in the form:
1
2

r 2 u2 (cp T'2 + u2 u'2 ) = Q' cp (T02 T01 )(r 1 u'1 + r'1 u1 )+r 1 u1 (cp T'1 + u1 u'1 ).

(3.8)

After using equation (3.4d) to expand cp T'2 and taking the limit u :0, this simplifies to
{r 2 u2 c22/(g 1)cp }S eiv(t b/u2 ) = Q' cp (T2 T1 )r 1 u'1 .

(3.9)

It then follows from a combination of equations (3.4c) and (3.9) that, for low mean flow
Mach numbers,
(g 1)Q' (g 1)cp (T2 T1 )r 1 u'1
(g 1)Q'
+
=
+ (r 1 r 2 )u'1 ,
c22
c22
c22

u2 r'2 =

(3.10)

. .

570

since (g 1)cp (T2 T1 )/c = (T2 T1 )/T2 = 1 r 2 /r 1 . Finally, substitution for u2 r 2' in the
equation of mass conservation (3.5) leads to
2
2

r 2 u'1 = r 2 u'2 (g 1)Q'/c22.

(3.11)

The no flow jump condition (2.8) then follows directly from equation (3.11). The pressure
continuity condition (2.7a) can be deduced in a straightforward way from the low Mach
number limit of equation (3.6).
When there is a mean flow, the strengths of the acoustic and entropy waves are coupled
by the continuity conditions (3.5)(3.7) and need to be solved for together. For an open
end the downstream boundary condition has a particularly simple form (see equation
(3.2)), that just relates the strengths of the upstream and downstream propagating acoustic
waves. For more complex geometries when, for example, the duct is terminated by a
nozzle, the entropy and acoustic waves may also be coupled by the exit boundary
condition.
A description of how the unsteady flow perturbs the rate of heat input is needed before
the frequency of the thermoacoustic oscillation can be determined. As in section 2, two
particular forms for the relationship between the rate of heat input and the flow will be
considered in detail. The first case is that of no unsteady rate of heat input;
Case I,

Q'(t) = 0,

no unsteady rate of heat input.

(3.12)

The second case is that of no unsteady heat per unit mass, the constant heat input per
mass being that required to raise the mean stagnation temperature from T01 to T02 :
Case II, Q'(t) = cp (T02 T01 )(r 1 u'1 + r'1 u1 ),

no unsteady heat input per unit mass.


(3.13)

Substitution for the unsteady flow from equations (3.3) and (3.4) into the boundary
conditions (equations (3.1) and (3.2)), the jump conditions (equations (3.5)(3.7)) and the
heat release equation (either equation (3.12) or (3.13)) leads to six homogeneous equations
for six unknowns. These equations can be written in the form
A
F
B
G
G
C
XG
D
G
2
ivb/u
G S(r2 c2/cp ) e
f
Q
/c1

J
G
G
G = 0,
G
G
j

(3.14)

where X is a 6 6 matrix.
For Case I,
1)
(1 M
0
F 1 + M 1
G (1 + M 1 )e1 (1 M 1 )e2 (1 + M 2 )e3 c1 /c2
G (1 + M 1 )2e1 (1 M 1 )2e2
 2 )2e3
(1 + M
X =G
G(1 + M 1 )a1 e1 (1 M 1 )a2 e2 (1 + M 2 )a3 e3 c2 /c1
0
0
e5
G
f
0
0
0

(1 M
 2 )e4 c1 /c2

M
 2 c1 /c2

(1 M
 2 )2e4

M
 22

(1 M
 2 )a4 e4 c2 /c1

e6
0

1
2

M
 32c2 /c1

0
0

0J
0G

G
G
1G
0G
1j
0

(3.15)

571

with
e1 = eivb/c1 (1 + M 1 ),

e2 = eivb/c1 (1 M 1 ),

e3 = eivb/c2 (1 + M 2 ),

e4 = eivb/c2 (1 M 2 ),

e5 = eivl/c2 (1 + M 2 ),

e6 = eivl/c2 (1 M 2 ),

a1 = M
 1 + 12 M
 21 + (g 1)1,
a3 = M
 2 + 12 M
 22 + (g 1)1

a2 = M
 1 12 M
 21 (g 1)1,
and

a4 = M
 2 12 M
 22 (g 1)1.

For Case II, with no unsteady heat input per unit mass, the last row in the matrix
defining X is to be replaced by
cp (T02 T01 )(1 + M
 1 )e1 /c12

cp (T02 T01 )(1 M


 1 )e2 /c12

1.

(3.16)

Equation (3.14) has a unique trivial solution A = B = C = D = S = Q


= 0 unless the
determinant of X vanishes. Therefore only disturbances for which det X = 0 can
propagate. For given mean flow properties, this is an equation for v and, in general, it
has complex roots. Since the time dependence is of the form eivt, the sign of Imaginary
v determines whether disturbances grow or decay. Real v gives the frequency of the mode.
It is a straightforward matter to determine the zeros of det X numerically, and some
results are shown in Figure 5. Inspection of the form of the matrix X in equation (3.15)
shows that the leading order effect of the mean flow is O(M
). However, it is apparent from
the numerical results that the importance of the the mean flow depends not only on the
inlet Mach number and stagnation temperature ratio (which affect the value of M
 2 ), but
also on the form of the coupling between the heat input and the unsteady flow. It is striking
how significant Mach number effects can be. For example, one might be tempted to think
that an inlet Mach number of 015 is sufficiently small for the mean flow to be
neglected. But at this Mach number, the frequency of the thermoacoustic oscillation for
Case II, T02 /T01 = 6, is reduced to half its no flow value. Care needs to be taken before
simply concluding on the basis of a low inlet Mach number that the mean flow may be
neglected.

Figure 5. The lowest frequency of oscillation as a function of inlet Mach number for b = l/2. T
 01 = 288 K,
p2 = 1 bar and T
 02 /T
 01 = 6. , Case I (no unsteady rate of heat input); , Case II (no unsteady heat input
per unit mass); W, Case I with no mean flow (root of equation (2.12)); Q, Case II with no mean flow (root of
equation (2.14)).

. .

572

4. THE EFFECTS OF DRAG

Most practical ways of adding heat also exert a drag force on the flow. In this section,
the effect of this drag on the frequency of thermoacoustic oscillations is investigated.
If the instantaneous drag can be modelled quasi-steadily, the drag force per unit duct
area may be written in the form 12 CD r1 u12(t), where CD is a drag coefflcient. In particular,
when a bluff body or grid is used for heat transfer or to stabilize a flame, its blockage
reduces the duct area available for the flow. If the expansion from this reduced area is
abrupt, an elementary one-dimensional quasi-steady analysis can be used to relate the drag
coefficient to the blockage ratio r. This gives [12]
CD = (1 (1 r)1 )2.

(4.1)

For linear perturbations the momentum equation (3.6) becomes


p'1 + r'1 u12 + 2r 1 u1 u'1 = p'2 + r'2 u22 + 2r 2 u2 u'2 + 12 CD r'1 u12 + CD r 1 u1 u'1 ,

(4.2)

and the third row of the matrix X in equation (3.15) is replaced by


(1 + 2M
1s + M
 21s)e1

(1 2M
1s + M
 21s)e2

(1 + M
 2 )2e3

(1 M
 2 ) 2e 4

M
 22

0, (4.3)

where
s = 1 12 CD = 12 + (1 r)1 12(1 r)2,

(4.4)

after substitution for CD from equation (4.1).


Results for various inlet Mach numbers and blockage ratios are shown in Figure 6. The
drag force exerted by a grid or flame holder with a blockage ratio of 25% or less is found
to have a negligible effect on the frequency of thermoacoustic oscillations for the range
of inlet Mach numbers considered. A blockage ratio of 50% alters the frequency for Mach
numbers greater than 01.
5. THE EFFECTS OF DISTRIBUTED HEAT INPUT

So far, the heat input has been considered as concentrated at a single plane, but in many
practical applications it is distributed over a significant length of the resonator. For

Figure 6. The lowest frequency of oscillation as a function of inlet Mach number for b = l/2. T01 = 288 K,
p2 = 1 bar and T02 /T01 = 6 and various values of the blockage ratio. , No drag; , 25% blockage ratio;
, 50% blockage ratio.

573

example, when a flame is stabilized in a duct in the wake of a bluff body, the combustion
is initiated as the flow passes the flame holder, but the fluid continues to burn throughout
the downstream portion of the duct [4]. The conditions under which a distributed heat
input can be considered as concentrated are discussed in this section.
It is sometimes said that heat input with axial extent d can be treated as lumped at a
single plane provided d is small in comparison with the acoustic wavelength (or more
precisely provided d W c /v). However, this is an oversimplification when there is a mean
flow. Then there is also an entropy wave, with the much shorter length scale u /v (see
equation (3.4)).
In a region of distributed heat input the strengths of the acoustic and entropy waves
change continuously with position. It is then easier to determine the flow by working
directly from the equations of motion. The one-dimensional equations of mass, momentum
and energy conservation for the mean flow are respectively:
d
(r u ) = 0,
dx

d
(p + r u 2) = 0,
dx

d
q (x)
(c T + 12 u 2) =
,
dx p
r u

(5.15.3)

where q(x, t) is the rate of heat input per unit volume. When q (x) is specified as a function
of x, and the flow at one axial position is known, equations (5.1)(5.3) can be integrated
with respect to x in a straightforward way to determine the mean flow throughout the duct.
Linear perturbations can be calculated in a similar way. For oscillations proportional
to eivt, the unsteady one-dimensional equations of mass, momentum and energy may be
written in the form
(d/dx)(r u' + r'u ) = ivr',

(5.4)

(d/dx)(p' + r'u + 2r uu') = iv(r u' + r'u ),


2

(5.5)

(d/dx)[(r u' + r'u )(cp T + u ) + r u (cp T' + uu')]=q' iv[r'(cv T + u ) + r (cv T' + uu')].
(5.6)
1
2

1
2

The upstream boundary conditions determine the inlet flow perturbations. Then, once the
mean flow, the frequency v and the relationship between q' and the unsteady flow has been
specified, equations (5.4)(5.6) can be integrated along the duct, thus determining the flow
perturbations at all positions. The details are given in reference [5].
At a general value of v, the exit boundary condition is not satisfied. It is therefore
necessary to iterate in v, at each state calculating the flow in the duct, until the complex
values of v, for which the exit boundary condition is met, are determined. These are the
frequencies of the thermoacoustic oscillations. Only disturbances with these particular
frequencies satisfy all the boundary conditions and can exist as free modes of the heated
duct.
As an illustrative sample, consider the duct geometry of section 3, where the inlet flow
is isentropic and choked, and the exit is open. The inlet boundary condition is described
by equation (3.1), together with the isentropic relationships r' = p'/c12 and cp T' = p'/r 1 .
Equation (3.2) shows the exit boundary condition to be p'(l, t) = 0. The mean heat input
is uniformly distributed over a length d, centred on x = b:
q (x) =

cp r 1 u1 (T02 T01 )/d


0

for =x b= Q 12 d
.
for =x b= q 12 d.

(5.7)

The lowest frequency of oscillation has been calculated for this distributed heat input by
integrating equations (5.1)(5.6), for Case I:
q'(x, t) = 0,

no unsteady rate of heat input per unit volume.

(5.8)

. .

574

Figure 7. The lowest frequency of oscillation as a function of stagnation temperature ratio for a uniform
distribution of mean heat input over a distance d. Case I, no unsteady rate of heat input, q' = 0, with M
1 = 01,
b = l/2, T01 = 288 K, p2 = 1 bar. , Concentrated mean heat input; , d=005l; , d = 025l.

The frequencies predicted for two values of d/l are shown in Figure 7. Also plotted for
comparison
are
results
for
the
concentrated
mean
heat
input
q (x) = cp r 1 u1 (T02 T01 )d(x b). These are the roots of det X = 0, where X is the 6 6
matrix in equation (3.15). It is apparent from these plots that there are significant
differences between the predicted frequencies for concentrated and distributed mean heat
input. For example, distributing the mean heat over a length as short as 5% of the duct
can lead to a 25% change in the frequency of the oscillation at the higher temperature
ratios, even though vd/c1 is only 01.
The detailed form of the axial distribution of mean heat input is not so important. In
Figure 8, results are given for a triangular distribution in the rate of mean input:

4cp r 1 u1 (T02 T01 )(x + 12 d b)/d


q (x) = 4cp r 1 u1 (T02 T01 )(b + 12 d x)/d
0

for b 12 d Q x E b
for b E x Q b + 12 d .
for =x b= q 12 d.

(5.9)

The plots are virtually indistinguishable from those in Figure 7, where the mean input is
distributed uniformly over the same axial extent d and with the same centroid b.

Figure 8. As Figure 7, but with a triangular distribution of mean heat input.

575

Hence, even if the mean heat has only a modest spatial distribution over a length
d E 01c /v, it can still lead to appreciably different frequencies of oscillation from that
when the heat input is concentrated at a single axial plane.
These numerical results can be interpreted and explained analytically. In the Appendix,
the conservation equations (5.4)(5.6) are integrated across a region of heat input. There
it is shown (see equation (A13)) that the pressure and velocity fluctuations on either side
of a region of heat input are unaffected by its spatial distribution over a length d provided
vd
W1
c1

v
cp

and

r (u u2 )s' dx W p'.

(5.10)

Here s' denotes the perturbation in entropy and 1 and 2 are the positions x = b 12 d and
b + 12 d on either side of the region of heat input. Entropy fluctuations, with their shorter
length scale u /v are more strongly influenced by a distribution in heat input. If the entropy
waves propagating downstream of a region of heat input of axial length d are to have the
same strength as those produced by concentrated heat addition, it is evident from equation
(A15) that d must be small in comparison with u /v. For a low Mach number mean flow
this is a strong constraint.
In the examples considered in this paper, the pipe exit is open and the boundary
condition of zero pressure perturbation does not involve the entropy waves. Hence, the
two conditions (5.10) are sufficient to ensure that the frequency of oscillation is unaffected
by a spatial distribution of heat input. The entropy fluctuations in the region of heat input
must be determined before the integral in conditions (5.10) can be estimated.
For a mean flow and linear perturbations, the entropy equation rT Ds/Dt = q(x, t) may
be written in the form

1s' qR q' u' p'


1s'

.
+ u
=
q
u
p
1t
1x
p

(5.11)

Consider Case I, in which q'(x, t) = 0. For disturbances of frequency v and low mean flow
Mach numbers, equation (5.11) simplifies to
ivs' + u 1s'/1x = qRu'/pu .

(5.12)

When vd/u is very small in comparison with unity, u 1s'/1x is the largest term on the left
side of equation (5.12). After substitution for q (x) from equation (5.7), integration leads
to

00

us'(x) 0 O cp

T2
x b + 12 d
1
u'1
d
T1

for

vd
W 1,
u

(5.13)

where the initial condition s'1 = 0 has been used. Note that the estimate of u2 s'2 obtained
by putting x = b + 12 d in expression (5.13) agrees with the value calculated in equation (3.9)
for d = 0. Expression (5.13) may be used to estimate the integral in conditions (5.10) and
gives
v
cp

r (u u2 )s' dx 0 O

0 0

1 1

vd T2
1 p'
c1 T1
2

for

vd
W 1.
u

(5.14)

Hence the two conditions (vd/c1 )(T2 /T1 1)2 W 1 and vd/u W 1 are sufficient to ensure
that the frequency of oscillation is not altered by the distribution of heat input over a small
length d.

. .

576

When the heat input is caused by combustion, the mean flow velocities are often so low
that the limit vd/u w 1 is more appropriate. Then ivs' is the largest term on the left side
of equation (5.12) and substitution for q (x) from equation (5.7) leads to the estimate
ivs'2 0 O

00

11

cp T2
1 u'1 ,
d T1

(5.15)

and hence
v
cp

0 0

r (u u2 )s' dx 0 O r 1 u1

1 1

T2
1 u'1 .
T1
2

(5.16)

It then follows from the inequalities in (5.10) that the conditions for the frequency of
oscillation to be unaffected by a distribution in heat input are
vd
W1
c1

and

M
1

T2
1
T1

W 1,

with

vd
w 1.
u

(5.17)

It is now possible to explain the numerical results in Figure 7. For d = 025l, vd/c1 is not
small and the frequency of oscillation is substantially changed by the axial distribution in
heat input. For d = 005l, vd/c1 is much smaller than unity but vd/u1 is not. It is then clear
from expression (5.17) that for temperature ratios near unity the effects of the distribution
should not be significant but that it becomes important at higher temperature ratios. These
predictions are confirmed by the numerical results shown in Figure 7.
The second of conditions (5.17) implies that the sensitivity of the frequency of oscillation
to the axial extent of the heat input is reduced at low Mach numbers. That is confirmed
by the results in Figure 9 for M
 1 = 001. A comparison with Figure 7 shows that the
distribution of heat input does indeed have less effect on the frequency of oscillation at
very low Mach numbers.
It is evident from equation (5.11) that the particular form of unsteady heat input for
which
q'/q = (u'/u ) + (p'/p )
(5.18)
leads to s'(x, t) 0 0. When there are no entropy waves the conditions (5.10) reduce to the
simple requirement that d be acoustically compact.
The frequencies of thermoacoustic oscillations for this form of unsteady heat input are
shown in Figure 10. For distributed heat input, the frequency was calculated by integrating

Figure 9. As Figure 7, but with M


1 = 001.

577

Figure 10. As Figure 7, but for q'/q = (u'/u ) + (p'/p ).

equations (5.1)(5.6), with the unsteady heat input (5.18). Results for the concentrated
mean heat input were found by replacing the last row in the matrix X in equation (3.15)
by
cp (T02 T01 )(1 + gM
 1 )e1 /c12

cp (T02 T01 )(1 gM


 1 )e2 /c12

1,

(5.19)

and finding the roots of det X = 0. As predicted, for the form of unsteady heat input (5.18),
the frequency of oscillation is less affected by the distribution of heat input than that for
Case I with q' = 0 shown in Figure 7.
The results for Case II, in which the rate of release is proportional to the mass flow rate,
show similar trends. For a concentrated heat input, the frequency of oscillation is a root
of det X = 0, where X is described by equation (3.16). When the heat input is distributed,
the unsteady heat input generalises to
Case II,

q' u' r'


= + ,
q
u
r

no unsteady heat input per unit mass.

(5.20)

This is close in form to equation (5.18) and only weak entropy waves are generated. Once
again, the frequency of oscillation is altered little by the distribution in heat input.
In summary, entropy waves have a short length scale, u /v, and are more affected by
an axial distribution in the heat input than the acoustic waves with their longer length scale

Figure 11. As Figure 7, but for Case II with q'/q = (u'/u ) + (r'/r ).

578

. .

c /v. Acoustic and entropy waves are coupled through a region of heat input. The two
conditions vd/c1 W 1 and v f r (u u2 )s' dx/cp W p' must be satisfied, if the acoustic waves
are to be unaffected by the distribution of heat input over an axial distance d. Results in
Figure 7, for M
 1 = 01 and Case I with no unsteady rate of heat input per unit volume,
show that distributing the heat input over a length as short as 5% of the duct can lead
to a 25% change in the frequency of oscillation at the higher temperature ratios. To
guarantee in general that heat input distributed over a length d can be treated as
concentrated, one requires both vd/u1 W 1 and vd/c1 W 1. This is very restrictive.
However, flow conditions that reduce the volume term f r (u u2 )s' dx, that accounts
for the coupling between the acoustic and the entropy waves, reduce the sensitivity of the
thermoacoustic oscillations to the axial extent of the heat input. The results in Figures 711
show that this can be achieved in several ways:
(i) A reduction in the mean temperature ratio. The results for small (T02 T01 )/T01 in
Figure 7 show that then the frequency of oscillation is only slightly affected by a modest
distribution in heat input.
(ii) A reduction in mean inlet Mach number. A comparison of Figures 7 and 9 shows
that this reduces the effect of the axial extent of the heat input.
(iii) Particular forms for the unsteady heat input. The unsteady heat input in equation
(5.18) generates no entropy waves, while only weak waves are produced when the heat
input per unit mass is constant. The results in Figures 10 and 11 show that then the
frequencies of the thermoacoustic oscillations are insensitive to a spatial distribution in
heat input, when the downstream boundary condition involves only the acoustic waves.
6. CONCLUSIONS

Model problems with very simple geometries have been considered to investigate the
influence of various flow effects on the frequency of thermoacoustic oscillations.
The form of the coupling between the heat input and the unsteady flow has been
demonstrated to have a crucial effect on the frequency of oscillation. Indeed, for the
particular case of heating at a single axial plane in a duct with one open and one closed
end, there is nearly a 60% difference between the frequency predicted for no unsteady heat
input per unit mass and no unsteady rate of heat input. A number of calculation methods
recommended in the literature have been tested by applying them to model problems. This
has shown that they do not account fully for this effect.
The presence of a mean flow significantly complicates the analysis of a thermoacoustic
oscillation, since an entropy wave becomes coupled to the acoustic field. It is, therefore,
tempting to neglect the mean flow, whenever the inlet Mach number is low. However, great
caution must be exercised before making such an assumption. Mean flow effects are found
to be surprisingly significant. For example, with a mean stagnation temperature ratio of
six, the frequency of a thermoacoustic oscillation for an inlet Mach number of 015 can
be reduced to half its no flow value.
Many practical ways of adding heat also present a blockage to the flow and exert a drag
force on the fluid. A drag force exerted by a grid or flame holder with a blockage ratio
of 25% or less is found to have a negligible effect on the frequency of thermoacoustic
oscillations for inlet Mach numbers in the range 0 E M
 1 E 015. A blockage ratio of 50%
alters the frequency for inlet Mach numbers greater than 01.
Even a modest distribution of the heat input over an axial distance d can lead to a
significantly different frequency of oscillation from that when the heat input is concentrated. For example, even an axial extent of heat input as short as 5% of the duct length
can lead to a 25% change in the frequency of oscillation. Entropy waves have a short

579

length scale u /v and are more affected by an axial distribution in the heat input than the
acoustic waves with their longer length scale c /v.
Hence, to guarantee, in general, the heat input distributed over a length d can be treated
as concentrated, one requires both vd/u W 1 and vd/c W 1. However, these conditions can
be relaxed when the downstream boundary condition involves only acoustic waves, and
the acoustic waves (and hence the frequency of oscillation) are not affected by the entropy
waves. This is the case at low mean flow Mach numbers or for mean temperature ratios
near unity. It also occurs for certain forms of the unsteady heat input, which lead to no
or only very weak entropy waves. In all these cases heat input, distributed over a distance
d, can be considered as concentrated provided that d is compact.
REFERENCES
1. L R 1896 The Theory of Sound, London: Macmillan.
2. S. M. C and T. J. P 1988 Proceedings of the Institute of Acoustics 10, 103153.
Interactions between acoustics and combustion.
3. F. E. C. C 1988 AGARD-CP-450. Combustion instabilities in liquid-fuelled propulsion
systemsan overview.
4. P. J. L 1988 Journal of Fluid Mechanics 193, 417443. Reheat buzzan acoustically
coupled combustion instability, part I: experiment.
5. G. J. B, A. P. D and P. J. L 1988 Journal of Fluid Mechanics 193,
445473. Reheat buzz: an acoustically coupled combustion instability, part 2: theory.
6. U. G. H, D. R, B. T. Z and B. R. D 1987 AIAA-87-0216. Fluid mechanically
coupled combustion-instabilities in ramjet combustors.
7. A. P. D 1988 AGARD-CP-450. Reheat buzzan acoustically coupled combustion
instability.
8. F. A. W 1965 Combustion Theory. Reading, Massachusetts: Addison-Wesley.
9. U. G. H, D. R and B. T. Z 1988 American Institute of Aeronautics and Astronautics
Journal 26, 532537. Sound generation by ducted flames.
10. A. P. D and J. E. F W 1983 Sound and Sources of Sound. Chichester: Ellis
Horwood.
11. A. M. C 1982 Journal of Fluid Mechanics 121, 59105. Low frequency sound radiation
and generation due to the interaction of unsteady flow with a jet pipe.
12. A. P. D and G. J. B 1984 AIAA-84-2321. Reheat buzzan acoustically driven
combustion instability.
APPENDIX

The aim of this Appendix is to determine conditions under which equations (5.4)(5.6),
describing changes through a distributed region of heat input, integrate to give the jump
conditions (3.5)(3.7) for concentrated heat input.
Integration of equations (5.4)(5.6) across the region of heat input, from position 1,
x = b = 12 d, to 2, x = b + 12 d, leads to

g
g
g

[r u' + r'u ]21 = iv

r' dx,

(A1)

(r u' + r'u ) dx,

(A2)

[p' + r'u 2 + 2r uu']21 = iv

[(r u' + r'u )cp T0 + r u (cp T' + uu')]21 Q'=iv

(r'(cv T + 12 u 2) + r (cv T' + uu')) dx,

(A3)
where Q' = f q'(x) dx. It is evident that equations (A1)(A3) reduce to the concentrated
heat input jump conditions when their right sides are negligible.
2
1

. .

580

In a perfect gas, the entropy s = cv ln p cp ln r + constant. For linear perturbations,


this may be rearranged to show that
r' = (p'/c 2) (r /cp )s'.
(A4)
Similarly,
cp T' = (p'/r ) + Ts'.
(A5)
The inlet boundary conditions ensure that there are no incoming entropy waves and so
s'1 = 0. Equations (A1)(A3) will now be rearranged so that they relate p'2 , u'2 and s'2 to
the inlet flow with additional integrals over the region of heat input.
First note that r'2 can be eliminated by subtracting the product of equation (A1) and
u2 from equation (A2) to give
p'2 + r 2 u2 u'2 = p'1 + r'1 u1 (u1 u2 ) + r 1 (2u1 u2 )u'1 iv

(r'(u u2 ) + r u') dx.

(A6)

A second equation for p'2 and u'2 is derived by noting that the energy equation,
[ru(cp T + 12 u 2)]21 = Q

1
(r(cv T + 12 u 2)) dx,
1t

(A7)

may be rewritten in the form

g
pu + 12 ru 3
g1

g 0
2

=Q
1

since cp rT = gp/(g 1). Equation (A8) is equivalent to

g
pu
g1

1
p
+ 1 ru 2 dx,
1t g 1 2

+ r1 u1 [12 u 2]21 + [ru]21 12 u22 = Q


1

g 0

(A8)

1
p
+ 1 ru 2 dx.
1t g 1 2

(A9)

For linear perturbations of frequency v this leads to


g
(u' p + u2 p'2 ) + r 1 u1 u2 u'2
g1 2 2
g
=Q' +
(u' p + u1 p'1 ) + r 1 u12u'1 + (r 1 u'1 + r'1 u1 )12 (u12 u22)
g1 1 1

g0
2

iv

p'
+ r uu' + r'12 (u 2 u22 ) dx,
g1

(A10)

after substitution for [r u' + r'u ]21 from equation (A1).


When r' is expanded in terms of pressure and entropy from equation (A4), equations
(A6) and (A10) become, respectively,
p'2 + r 2 u2 u'2 = p'1 + r'1 u1 (u1 u2 ) + r 1 (2u1 u2 )u'1

g 00
2

iv

p' r s'

(u u2 ) + r u' dx,
c 2
cp

(A11)

and
g
(u' p + u2 p'2 ) + r 1 u1 u2 u'2
g1 2 2
g
=Q' +
(u' p + u1 p'1 ) + r 1 u12u'1 + (r 1 u'1 + r'1 u1 )12 (u12 u22)
g1 1 1

g0
2

iv

p'
p' r
+ r uu' + 2 s' 12 (u 2 u22) dx.
g1
c
cp

(A12)

581

Equations (A11) and (A12) describe the changes in p' and u' across a region of heat
distributed over a length d. It is evident that these are equal to the changes for a
concentrated heat input provided that the contributions from the integrals are negligible:
i.e., provided that
vd
W1
c1

and

v
cp

r (u u2 )s' dx W p',

(A13)

and the flow is subsonic. These conditions are sufficient to ensure that the acoustic waves
on either side of a region of heat input are unaffected by its distribution over a finite length.
To determine s'2 , the strength of the entropy wave, note that equations (A1) and (A3)
may be combined to give a modified energy equation:
r 1 u1 (cp T'2 + u2 u'2 cp T'1 u1 u'1 ) + (r 1 u'1 + r'1 u1 )cp (T02 T01 ) Q'

g0

=iv

r'(cp T02 cv T 12 u 2) r (cv T' + uu') dx.

(A14)

Substitution for r' and T' from equations (A4) and (A5) into equation (A14) leads to

r 1 u1 T2 s'2 +

g0
2

+iv

p'
p'2
+ u2 u'2 = Q' + r 1 u1 1 + u1 u'1 (r 1 u'1 + r'1 u1 )cp (T02 T01 )
r 2
r 1

s'

r
p'
(c T 1 u 2) + 2 cp (T02 T0 ) r uu' dx.
cp p 02 2
c

(A.15)

Estimation of the integral in equation (A15) shows that


vd/u W 1

(A16)

is required if the strength of an entropy wave produced by heat addition is to be unaffected


when the heat input is distributed over a length d.
The implications of the conditions (A13) and (A16) are discussed in section 5.

Вам также может понравиться