You are on page 1of 25

Journal of Applied Mathematicsand Physics(ZAMP) 0044-2275/85/002250-25

$ 6.50/0
Vol. 36, March 1985 9 BirkhS.userVerlagBasel,1985

Thermally induced low-frequency oscillations


By J. J. Keller, W. Egli and J. Hellat, Brown Boveri Research Center,
CH-5405 Baden, Switzerland

1. Introduction

In recent years, substantial efforts have been made to understand the phys-
ical mechanisms which lead to combustion oscillations. The first comprehensive
investigation of combustion instabilities in rocket motors was made by Crocco
and Cheng [1], who distinguished between low-frequency instabilities (chugging)
and high-frequency instabilities (screaming). In the former case typical acoustic
wavelengths are very long compared to the dimensions of the combustion cham-
ber, such that the flow in the combustion chamber can be considered as quasi
steady, whereas in the latter typical acoustic wavelengths are comparable to the
combustion chamber dimensions. Subsequent investigations of unstable com-
bustion in rocket motors have been concerned mainly with screaming. Based
upon the sensitive time-lag model devised by Crocco, a second-order theory for
screaming (nonlinear longitudinal instability) was developed by Mitchell,
Crocco and Sirignano [2]. So far as low-frequency oscillations are concerned,
however, no comparably comprehensive investigation has been made. It is im-
portant to recognize a fundamental difference between the typical low-frequency
behaviour of rocket motors and that of industrial combustion chambers and
furnaces. In a rocket combustor with concentrated combustion and an acousti-
cally compact exit nozzle, both the combustion zone and also the nozzle
entrance essentially represent acoustically closed ends, whilst the inlets and
outlets of industrial combustion chambers represent effecitively open ends. To
first order (i. e. in the low-frequency limit) the consequence of this difference is
that in the former case only the pressure oscillates, i.e. for low-frequency oscil-
lations the rocket combustor behaves essentially like the cavity of a Helmholtz
resonator. In the latter case only the particle velocity oscillates, which essentially
corresponds to the oscillation of the gas column in the throat of a Helmholtz
resonator. In the present paper we restrict consideration to longitudinal insta-
bilities of the latter kind. It is well known that industrial combustion chambers
often produce oscillations with frequencies considerably below those of the
fundamental acoustic modes and that these frequencies depend strongly on
temperature, throughflow, geometric dimensions and flame structure.
Vol. 36, 1985 Thermally induced low-frequency oscillations 251

Having specified the nature of the oscillations to be discussed, we proceed


to discuss the excitation mechanism considered. The model problem envisaged
herein is restricted to quasi-one dimensional flows. Furthermore, the dimensions
of the exit nozzle (or exit contraction) of the combustion chamber are assumed
to be small compared to the chamber length, and the cross-sectional area of the
chamber is assumed to be constant. As a first step we also assume that the flame
is short compared to the chamber length. Then, if the mixing of the reactants is
highly homogeneous, the combustion zone can be considered simply as a zone
of heat addition and the effects of mass and momentum addition owing to fuel
injection may usually be regarded as negligible, as will become evident in the
course of the analysis. In the simplest case the combustion zone could be
replaced by a short heater element. Having excluded the possibility of sound
generation due to chemical inhomogeneities and ignoring Reynolds-stress ef-
fects, i.e. turbulent fluctuations, the only remaining excitation mechanism is
that resulting from convection of gas containing zones of nonuniform tempera-
ture or "entropy spots". Since the chamber flow is assumed to be quasi-one
dimensional the problem is reduced to the consideration of "entropy slugs". The
generation of sound by convection of entropy nonuniformities has been investi-
gated previously by Candel [3], Marble [4], Cumpsty and Marble [5], Ffowcs
Williams and Howe [6], Marble and Candel [7], Keller [81 and others. As envi-
saged here, the problem involves a pressure disturbance which propagates up-
stream into the combustion zone near the inlet of the chamber where it is
scattered thereby producing a modulation of the equivalence ratio. The distur-
bed equivalence ratio then gives rise to entropy disturbances in the flow down-
stream of the combustion zone and these are convected downstream producing
pressure disturbances in the exit nozzle. The resonance problem thus defined will
be investigated on the basis of a second-order analysis taking into account
entropy nonuniformities. Following Mitchell et al. [2] we choose the mean-flow
Mach number (which is assumed to be small compared to unity) as the expan-
sion parameter. This particular choice provides a simplification which is essen-
tial for a successful second-order treatment of the problem. Assuming that the
entropy wavelength scales with the length L of the combustion chamber, a
typical sound wavelength is by an order 0(M -1) longer than L. On the other
hand we have assumed that combustion is limited to a discrete axial location.
Hence the dimensions of the combustion zone and the exit nozzle may be
assumed to be by 0(M) smaller than L. Thus we have a compact combustion
zone and exit nozzle even to second order. The consequence is that no time-lag
terms arise from the boundary conditions which may, therefore, be considered
as quasi steady. In practice the combustion zone often extends over a consider-
able part of a combustion chamber. For this reason modifications appearing due
to distributed combustion will be discussed towards the end of this report. A
criterion for the compactness of both the exit nozzle and the combustion zone
can be obtained from a wave equation derived by Keller [8, Eq. (26)]. Using
252 J.J. Keller, W. Egli and J. Hellat ZAMP

this equation it may be shown that in the low-frequency limit the following
first-order proportionality relations hold
(1 + M)f+ + (1 -- M ) f ,,~ Mco, (1.1)
(1 + M ) f + - ( 1 - M ) f + 2 z , , ~ C o 1, (1.2)
z ~ Co, (1.3)

where

= ca -- z)
, (1.4)

Co S1
z - (1.5)
(y - 1) %
and M, c 0, c 1, u j, sl, cp and 7 denote the mean-flow Mach number, the un-
disturbed soundspeed, the soundspeed disturbance, the particle-velocity distur-
bance, the specific-entropy disturbance, the specific heat per unit mass at con-
stant pressure and the ratio of specific heats, respectively. The suffix "1" refers
to first-order quantities.
It may be noted that Eqs. (1.1) to (1.3) imply the low-frequency scattering
relations derived by Marble and Candel [7]. If the term containing the angular
frequency o) in the wave equation of Keller, referred to above, is considered as
a small disturbance term and a length scale I is introduced for the nozzle (or in
general, a zone of nonuniform flow), it is straightforward to arrive at the follow-
ing compactness criterion

1, (1.6)
(1 - M ) c o

whereby M and c o should be taken at the nozzle entrance. It is surprising that


for small values of M Eq. (1.6) implies only that I should be small compared to the
sound wavelength and not necessarily small compared to the entropy wavelength.
Finally, before we proceed to the analysis, an important difference between
entropy slugs and entropy spots should be discussed briefly. Even with perfect
mixing in the combustion zone it is necessary to extend consideration to entropy
spots if, for instance, cold air is injected downstream of the combustion zone (e. g.
for film cooling of the chamber walls). On the basis of an investigation by Ffowcs
Williams and Howe [6], who considered the interaction of general entropy spots
with regions of nonuniform flow, a comparison can be made between the intensi-
ties of low-frequency sound waves produced by entropy slugs and the intensities
for more general entropy spots (e. g. a hot slug surrounded by a cold film). In the
latter case it is assumed that the enthalpy and mass flow averaged over any
cross-sectional area of the combustion chamber are equal to the enthalpy and
mass flow at the corresponding axial location in the former case. It then turns
Vol. 36, 1985 Thermallyinduced low-frequencyoscillations 253

out that in the case where the entropy is uniformly distributed over the cross-
sections (i. e. entropy slugs) the intensities of the sound-waves produced are the
smallest. However, the additional effects of cross-sectional entropy nonuniformi-
ties are usually small compared to the intensities of soundwaves produced by
entropy slugs.

2. Analysis

When heat conduction and viscous effects are ignored, the equations of
motion governing the one-dimensional flow between the zone of combustion
and the exit nozzle are
2 dc ~u
+ 0, (2.1)
7 -- 1 dt C~xx
du 2 ~c c ~S~
d t + 7 - 1 c ( ~xx 2 cp 8xJ = O, (2.2)

ds
dt 0, (2.3)

where
d ~D
- +
dt St u ~xx
and u, c and s denote particle velocity, soundspeed and specific entropy, respec-
tively. Based upon the discussion of wavelengths and dimensions in the Intro-
duction, it is postulated that equations (2.1) to (2.3) can be solved iteratively with
the help of the following expansion scheme

u = ~ M"u, + k M"U,, (2.4)


n=l n=3

C=Co+ M"c.+ M~176 12.51


n=2 n=0

s/cp = ~ M" Z,, (2.6)


n=0

wherein the components u, and c. represent the isentropic contributions to the


solutions and U,,, C, and Z , account for the modifications due to entropy
convection. The zeroth-order component c o is assumed to be the soundspeed
downstream of the zone of combustion in the undisturbed case. Although some-
what unusual, this splitting of the expansions turns out to be very convenient.
Note that in (2.4) and (2.5) the components U1, U2 and c 1 have been set
254 J.J. Keller, W. Egli and J. Hellat ZAMP

equal to zero. One of the crucial problems of the subsequent analysis is to show
that setting U 1 and U2 equal to zero is compatible with the orders of magnitude
of entropy and sound wavelengths specified in the Introduction. The first step
is now to determine the components u, and c,. As the inlet and exit of the
combustion chamber represent effectively open ends and the acoustic wave-
lengths are large compared to the chamber length it is clear that c I should indeed
be set equal to zero. Hence, after introducing the Riemann invariant f and
integrating the isentropic parts of (2.1) and (2.2) iteratively to second order, using
the usual integration technique (see e. g. Keller [9]), and expanding all terms with
respect to x, we obtain

M x 2 f,, M2 7 + 1 M2 f,
u = M c o + M f (t) + ~ c~~ (t) + - - x f ' (t) + - - x (t) f (t)
co 4c 2

+ MZx [f'(t)f(t) - F(t)f"(t)] + - - 2 - [ g ( t ) + h(t)]

x M e [g, (t) - h' (t)] + 0 (M 4 Co), (2.7)


2co2

2 x 3 -TM 2
- - c = -- M - - f '(t) + F(t) f'(t)
? - 1 Co ~oCo

M 2
+ ~ [g(t) -- h(t)] + 0 (M 4co) (2.8)

where M d e n o t e s the ratio of mean particle velocity to undisturbed soundspeed,


M . (co + f ) represents the first-order approximation for the particle velocity, g
and h'are functions of integration and F is defined by

F' (t) = f (t),


whereby the mean values of both f and F are assumed to be zero. These
assumptions are compatible with the definition of c o . Making use of the fact that
the angular fl'equency co is of the order

0 (co) = M c o / L , (2.9)
we obtain from a comparison of (2.4) and (2.5) with (2.7) and (2.8), respectively,

ul = Co + f (t),
1
= [o (t) + h (t)], (2.1o)

1
2 x f,(t)+ [g (t) -- h (t)] .
-- 1 c2 - M Co
Vol. 36, 1985 Thermallyinduced low-frequencyoscillations 255

Furthermore, we note that the third-order components u 3 and c 3 contain


secular terms in x. N o w we can proceed to determine the components U,, C, and
Z,. After introducing u t (according to (2.10)) in (2.3) we obtain, after integration,

x c M F (t)'] (2.11)
Z o = r/ t Mc o J'

Making use of the first law of thermodynamics, i.e.

ds dT 7 - I dp
cp T y p
where T and p denote temperature and pressure, introducing the expansions
(2.4)-(2.6) in (2.2) and neglecting third-order terms yields with the help of
(2.10).

[co+Co]2~p_ 2 [~c c ~sl_ 2 M2~C2 Mf,(t),


?p ~x 7 - 1 c ~x 2% ~xx 7 - 1 co ~x
(2.12)
which implies, for example, that to first order

[Co + MC~] = ~ {[co + Co] + MC1} ~xx[Zo + MZ1]. (2.13)

Thus we have, with the help of the mean-value conditions,

2 In 1 + = Zo. (2.14)

2C 1=[c o+Co]Z 1 (2.15)


and
~Co ~Z o
z l ~ - x - c l ~ - x = 0. (2.16)

From (2.14) and (2.15) it is easy to see that the zeroth- and first-order terms in
(2.1) and (2.3) are compatible only if U 2 is set equal to zero, as was done in the
expansion (2.4). In particular U 2 = 0 implies that (see Eq. (2.1))

0 C-~x = c2 M3/L. (2.17)

Hence, we obtain from (2.3) (or from (2.1))

M [6 (t) + H (t)] i/' t +M~ t


x --MF(t)']
z, = 2c-7 / Mco /'
(2.18)
256 J . J . Keller, W. Egli a n d J. Hellat ZAMP

where primes denote derivatives with respect to the argument, the mean values
of G and H, which are integrated functions of 9 and h, are assumed to be zero
and r is a function of integration. According to (2.17), Eqs. (2./) and (2.3) are no
longer equivalent when second-order terms are retained. It is straightforward
but tedious to show that the additional second-order terms in Eq. (2.1), which
are responsible for the difference with respect to (2.3), can be balanced by a
suitable choice of U3.
1
A crucial point is now that u 2 = g [9 + hi and ~ can be set equal to zero after
suitably redefining the solution functions f and t/, i.e.
1
f , ew = fo~a + ~ M - [9 + h], r/,ew = ~/ola + M 4-
This redefinition corresponds to a first-order transformation which has been
discussed by Keller [9]. It does not affect the first approximations of f and ~/.
Thus we have from (2.5), (2.11), (2.12), (2.14) and (2.15), making use of Z1 -= 0 and
U 2 ~ 0,

1 ~p ~Mf'(t)
- + 0 (M4/L) (2.19)
p ~x [Co + Co] 2
and
c = Co + Co/It x - MMc oF (0 /J + 0 (M 2 Co). (2.20)

It should be pointed out that terms of order 0 (M3/L) and 0 (M Co) do not appear
in expressions (2.19) and (2.20), respectively.
It is remarkable that in the course of deriving the explicit expression (2.19)
for the pressure gradient it was not necessary to assume that the soundspeed
disturbance C Oshould be small. The expressions (2.19) and (2.20) are valid even
if 0 (Co) = 0 (c), implying that the amplitudes of the temperature oscillations
may be comparable to the mean temperature.

3. The boundary conditions

We consider the model combustion champer depicted in Fig. 1. As was


assumed in the Introduction, we regard the zone of combustion and the exit

zone of M N
heat addition I I

air ~ I
fuel ~ IA

air x
I I =
0 L
Figure 1
Idealized m o d e l c o m b u s t i o n c h a m b e r with a c o n c e n t r a t e d zone o f c o m b u s t i o n at x = 0 a n d a
c o m p a c t nozzle at x = L.
Vol. 36, 1985 Thermally induced low-frequency oscillations 257

nozzle as short compared to the entire combustion chamber and the Mach
numbers to be everywhere small compared to unity. Furthermore, to keep the
downstream boundary condition as simple as possible, we assume that the exit
nozzle exhausts into a large settling chamber of constant pressure. In the subse-
quent discussion quantities at the nozzle entrance and exit are denoted by the
suffices "M" and "N".
Based upon the assumption of compactness, the nozzle flow can be consid-
ered as quasi steady (i. e. also quasi isentropic). When the contraction parameter
r, defined by
A uN
r ----- -- q- 0 ( m 2 ) , (3.1)
AN um
is introduced in the pressure ratio

PM_I+ 7 [U2 _ u 2 ] + 0 ( M 4) (3.2)


Pu 2 c~
we obtain to second order

PM -- Pu _ • [r2 _ I] . M 2 I
PU 2 CO + C o ( L , t )
Co+f(t)12 ' (3.3)
where PM = P (L, t).
To discuss the upstream boundary condition we introduce the suffices "C"
and "H" which refer to quantities before (cold) and after (hot) the zone of
combustion and for simplicity we assume that the zone of combustion can be
considered as an axial location (x = o) of constant heat addition uniformly
distributed over the chamber cross-section. To justify these simplifying assump-
tions it should be pointed out that for liquid fuels it is usually the case that the
rate of fuel addition does indeed respond only weakly to pressure oscillations in
the zone of combustion. For gaseous fuels on the other hand, the supply system
often plays a crucial role. Furthermore, the addition of mass, momentum and
energy (not including the chemical energy) due to fuel injection represents a
small effect compared to the corresponding mass flow and the momentum of the
air in front of the zone of combustion and the heat addition in the zone of
combustion. Although retaining these additional effects would not lead to signif-
icant difficulties, they have been ignored in the interests of simplicity. If we
denote the rate of heat addition per unit cross-section by Q and the Mach
number in front of the zone of combustion by M c = Uc/Cc, we obtain from the
one-dimensional, quasi-steady equations of motion

_un_ = 1 + ~ - 1 Q + 0 ( cM
), 2 (3.4)
Uc 7 UcPc
Tn
--=,
~CH12
, =1+
7--1 Q
+O(M~) (3.5)
Tc LCcA 7 UcPc
258 J.J. Keller, W. Egli and J. Hetlat ZAMP

and

P~=1-[7-1] Q M 2+0(M2). (3.6)


Pc UcPc
Furthermore, we define a pressure drop Ap across the inlet by
Ap ?
- 2~M~. (3.7)
Pc
Note that 2 = I would correspond to a loss-free inlet. Here it should be pointed
out that the possibility of flow reversal is excluded, i.e. we require u c > O.
Finally, the possibility of flame oscillations should be discussed briefly. It is
apparent that the zone where most of the heat release takes place oscillates back
and forth owing to the oscillating particle velocity u. However, as an oscillation
of the flame affects the present problem to second order only, this has no
influence on the stability limit. Furthermore, as long as the displacement of the
flame from its mean position is always small compared to L, its second-order
effects are also negligible. Again to keep the analysis simple, rather than to avoid
essential difficulties, we ignore oscillations in the flame location. From (3.4) and
(3.5) we have

[CO Jl- C 0 (0, t)] 2 = CO -{- f (t) -~ 0 (M2), (3.8)


L Cc J Co - ~ + f (t)
where
7-1Q
M ~= -- (3.9)
? Pc
and u u is identified by

u~r = u = M [c o + f ] + 0 (M 2 Co).

Note that

Cc = (Cc} + 0 (M~ Cc) ,

where ( } refers to the time average.


From (3.4), (3.6), (3.7) and (3.8) we obtain, to second order,

_ ? M z{c o_~+f(t)} ~+~[c o-~+f(t)] . (3A0)

Thus the problem is reduced to solving Eq. (2.19) for the boundary conditions
(3.3) and (3.10).
Vol. 36, 1985 Thermally induced low-frequency oscillations 259

4. T h e nonlinear w a v e equation and its stability limits

For convenience we define the following dimensionless quantities:


= M c o t/L,

q) (r) = f (O/c o ,

r ('0 = Co (O/Co, (4.1)


S = I - UCo,
R = (r 2 - 1)/2.
Note that S = T c / T o ( = T c / T u in the small-amplitude limit), where the definition
of To corresponds to that ofco. Combining (2.19), (3.3), (3.8) and (3.10) leads, with
the help of (4.1), to

(p' (,) 1 d0 + R. [ I + qo ( ' c ) ) ] 2


o [l + r (~ - o + o~ (~))]~ I + r (~ - l + 9 (~)
1 2
+ ~ [S + ~o('c)l {1 - S + ~ [S + q) ('c)]} = f2 (4.2)

where
9 ' (0 = ~o (~), ( ~ ) = 0,
(4.3)
I + e (-c)
[l + ~ (~ + 45 (~))12 = s
s + q~ (~)
and f2 is a constant of integration which depends on the solution qo. Physically
f2 represents the (dimensionless) pressure drop across the combustion chamber.
Hence g2 reaches its absolute m i n i m u m in the limit of small oscillation ampli-
tudes. This property is also obvious from Eq. (4.2). F u r t h e r m o r e we note that
(4.2) admits continuous solutions only. If ~0(-c)were discontinuous, the first term
in (4.2) would necessarily contain a Dirac delta function (note that the value of
the integral is strictly positive) which could not be balanced by any other term.
To determine the linear neutral stability limit we now assume that the
amplitude of ~0 is infinitesimally small. Linearizing (4.2) then leads to

qY (-c) + ~ - R q~(z -- 1) + + 2R + ~o('c) = 0. (4.4)

The domains of instability are obtained after introducing


qo(z) ~ exp ((2 rc i ~ + v) ~) (4.5)
with the restriction v > o. The stability limit (i. e. v = o) is given by
1-S
2 rc # - S R sin 2 u ~t (4.6)
260 J. J. Keller, W. Egli and J. Hellat ZAMP

and
1-S
(4.7)
S
The domains of the normalized frequency #, corresponding to values of R and
S in the intervals
I__<R<~,
0<S<1,

for which (4.6) and (4.7) can be satisfied are near (just below) the values
2n--1
where n is a positive integer.
2
In other words the entropy wavelength for the nth mode is somewhat longer
than 2 L / ( 2 n - 1). The asymptotes (which are the same for all modes)
S ~ 0, R ~ I (4.8)
and
R --+ o% S --+ 1/3 (4.9)
both correspond to the limiting values # = (2 n - 1)/2. Furthermore, when the
unstable domain is entered, the first harmonic appears first and stability limits
of higher modes are always reached in monotonic sequence. Figure 2 shows the
stability limits of the first two modes for the values 2 = I and 2 = 2.
Considering periodic solutions with arbitrary mean values the wave Eq. (4.2)
admits a scaling transformation which turns out to be very useful for the

~0

~. first mode

,/g
I I I I
0
.7 .8 .9 TN-Tc t.0
Figure 2 TH
Linear neutral stability limits for the first two modes. The solid lines c o r r e s p o n d to a loss-free inlet
(Z = 1), the b r o k e n lines illustrate the effect o f a small inlet loss (2 = 2).
Vol. 36, 1985 Thermally induced low-frequency oscillations 261

numerical c o m p u t a t i o n of periodic solutions. F o r convenience we introduce the


following definitions
g(z) = [1 + 0 ('c)]2 , (4.10)
m = (~o), i.e. the mean value of qo, (4.11)
fi -- 1 + m, (4.12)
s/~
K - (4.13)
~-~+s
The transformation can now be defined by

q) = flq~* + m ,
g = Kg*,
S = K S*, (4.14)
R = R*,
2 = ,~*,
f2 = fiz f 2 * / K .

To avoid complicated notation, arguments are not displayed. However,


the interpretation is always that untransformed functions depend on un-
transformed arguments and transformed functions depend on transformed
arguments. Hence "~0 = flop*+ m", for example, should be interpreted as
"~o(~) =/~ ~o* (z*) + m".
The following relations are obtained immediately from (4.14)

l+cp=(l+q0*)fi, ~=~*+-~-z , = cp , (4.15)

where the primes denote the derivatives with respect to the arguments. Introduc-
ing (4.14) and (4.15) in (4.2) the transformed wave equation becomes

<p*' (z*) [' dO + R* [1 + q)* (z*)] 2


a* (~* - 0 + ~* (~*)) g* (~* - 1 - ~* (<))

1 [S* +
+ ~u ~o, (r*)] { 1 - s, + ~~*- [ S * + q)* (z*)]} = ~2", (4.16)

where
qs*' = ~o*, (<p*) = 0
and
1 + q)* (~*)
g* (~* + ~b* (~*)) -- S* S* + cp* (~*)" 0.17)
262 J. J. Keller, W. Egli and J. Hellat ZAMP

According to (4.11) m is defined to be the mean value of ~0. It is for this reason
that the mean value of ~0" is zero. The transformation could of course be
extended to arbitrary constants m, i.e. to periodic solutions q)* with arbitrary
mean values. However, the restriction (4.11) has been included because the
transformation will be used later to produce solutions with zero mean value.
A further point to be discussed concerns the excitation mechanism of the
oscillations, which is defined by (4.3). Equation (4.3) essentially represents a
relation between velocity disturbance and temperature disturbance at the down-
stream side of the combustion zone. In Sect. 3 we have already excluded the
possibility of flow reversal. However to provide a reasonable model for the
combustion process, for which Eq. (4.2) does not admit singular solutions, a
saturation effect should be included. There should be an upper limit for the
temperature to account for the fact that decreasing the flow speed leads to a
corresponding increase of the downstream temperature only as long as the
equivalence ratio is larger than 1 (or more precisely, as long as there is sufficient
oxygen for a complete combustion process). The simplest extension of (4.3),
which accounts for such a saturation effect, is
+
g (z + r (z)) = rain + q~(z) (4.18)

[aS
where a is the largest possible value of the temperature ratio Tn/Tc. The possi-
bility that q) < - S is not considered. According to (4.18), in addition to the
linear neutral stability limit described by (4.6) and (4.7), there is the limit S = 1/a.
Smaller values of S are excluded by the definition (4.18).

5. The solution of the wave equation

5.1 An analytic approximation


As a first step we consider the asymptotic regime in which R is very large
compared to unity. In this case the first term in (4.2) can be neglected at least for
domains of the solutions where 0 (~o') =< 1. As will become obvious from the
numerical solutions presented subsequently, the domains where 0 (q)') > I are
represented by a small neighbourhood of the zeros of the solutions.
For large values of R there are two small domains per oscillation cycle
where rapid transitions from negative to positive (and vice versa) values of the
solutions occur. With the exception of these domains 0 ((p') < 1 is strictly satis-
fied. However, for the present consideration we strictly neglect the first term in
(4.2) and replace the rapid transitions by jumps. Thus it is possible to construct
approximate solutions represented by periodic step functions. For simplification
Vol. 36, 1985 Thermallyinduced low-frequencyoscillations 263

we consider the fundamental mode only. As R --, ~ the oscillation period


approaches the limiting value 2. Hence, we can define one period of the solution
~o in the interval 0 < 'c _< 2 by

~o('c) = { b a if0<'c<l+% (0<%<1) (5.1)


if 1 + Z o < ' c < 2 ,
whereby a discontinuous jump at 'c = 1 + % should be included. Making use of
the freedom to choose the location of the origin on the 'c-axis we have assumed
that the second jump occurs at 'c = 2. Integrating (5.1) yields

('c) =
[+i 1+% 2 ] if0<z<l+%
(5.2)
+%[z 3+% 1 if 1 + % < ' c < 2 .
a.] "Co 2
Combining now (4.18) and (5.2), together with the condition (see Eq. (4,2)) that
the discontinuities of q) (z) and ~ ('c - I + 9 (-c)) should coincide, leads to

S(t+%)a+(1_%)S

o('c) =
min I - a a
S --s--a if 1- (1+%)<'c<2+~(1+%),
where

% and b- "Co (5.4)


a- 1 +'co 1 --'co'
Inserting (5.1) to (5.4) in the wave Eq. (4.2) and ignoring the first term of (42)
yields two relations between % an f2 (one for each of the two domains where ~o
is constant). Eliminating Q finally leads to
R . {(1 - %)2 _ (1 + "co)/a} - 2 2 %
R.{(1-%)3-(1 +%)3}+2(1-2)(1-'co 2)%
if 1 > [ 1 - ( 1 + % ) ( 1 - S ) ] a
1 --S ~ (5.5)
R - {(1 -- %)2 _ (1 + % ) 2 } _ 2';+ Zo
R . {(1 - - "/70)3 - - (1 q- To) 3} + 2 (1 - 2) (1 - 'c~) %
if 1 < [ 1 - ( 1 + % ) ( 1 - S ) ] a .
For the special case (R = 40, 2 = 37, ~r = 7), which approximately corresponds
to the conditions for which the subsequent experimental results were obtained,
264 J. J. Keller, W. Egli and J. Hellat ZAMP
0.30

Figure 3
Peak-to-peak amplitude a + b = d, symmetry depar-
0.05~
tures % and b - a versus relative temperature in-
crease 1 - S. 0.75 0.80 ~ I-S 0.85

Fig. 3 illustrates %, b - a and a + b as functions of 1 - S. In subsequent sec-


tions the peak-to-peak amplitude a + b will be denoted by d.
The high-temperature stability limit Zo = 0 for the first solution branch
defined by (5.5) corresponds to S = I/a. The high-temperature stability limit for
the second branch (5.5) is (set Vo = 0)
2R+2
I - S - (5.6)
3R-1+2'
which agrees with the linear neutral stability limit S -- SL defined by (4.7) for
cos 2 ~ # = - 1.
The nonlinear low-temperature stability limit S -- SN corresponds to the
point of coincidence of the two branches (5.5), i.e.
[1 - (1 + % ) ( 1 - S)] o- - 1. (5.7)
Following the corresponding discussion by Mitchell et al. [2] of the stability
properties of the solutions, it appears to be the case that the first branch (5.5)
corresponds to stable solutions whereas the second branch represents a dividing
line between decaying and growing oscillations and therefore corresponds to
unstable solutions. It m a y be shown that
sN > sL. (5.8)
F o r this reason there is a hysteresis in the interval SL < S < SN. If the noise level
in the combustion chamber is very low S can be decreased (i. e., TH is increased)
almost d o w n to the linear neutral stability limit S = SL until an oscillation
appears. On the other h a n d if S is increased when a stable oscillation is present
it does not disappear until S exceeds the value S = SN. As the hysteresis interval
SL = 0.25 < S < SN = 0.25040 is very small for the case illustrated in Fig. 3, the
unstable solution branches, wich almost coincide with the second coordinate
axis, are not shown in Fig. 3.
Vol. 36, 1985 Thermallyinduced low-frequencyoscillations 265

5.2 Numerical solutions


Equations (4.2) and (4.18) can formally be written as a first-order system:

q0' (z) = F [~o, 4~; r, R, s, 2, s a], (5.9)


(+) = (+).

The numerical integration starts out from a small-amplitude sinusoid (with the
wavelength chosen according to the linear theory) as the initial function. The
aim is to find a suitable constant g2 for which the evolution produced by the
successive integration of (5.9) leads to a periodic solution ~0(~) asymptotically.
The numerical integration is carried out according to the Adams-Bashfort-
Moulton method (PECE). It should be pointed out that for the evaluation of the
functional F the convolution integral
1 dO
o0 - 0 +

can be computed recursively.


High accuracy was obtained with the choice of 200 integration steps per
linear period. The computation is restricted to fundamental modes. To obtain
a criterion for the growth or decrease of the "local amplitude" of ~0(t) in the
course of the evolution the amplitude norm
d = max (p (t) - min (p (t)
taken over the "local period" was introduced. The constant f2, which represents
the dimensionless pressure drop across the combustion chamber, was obtained
from the periodicity condition. Based upon the change Ad of the norm d over a
certain number of local periods ( ~ 4) it is possible to formulate the performance
function
P (s = (Ad) 2 .

The aim was to find a value (2 = s o for which P (s = 0. One of the difficulties
in obtaining a nontrivial periodic solution (0 (t) arose from the fact that Eq. (4.2)
admits two constants as solutions for every value of s Within a sufficiently small
neighbourhood of s the problem is convex. Hence it was possible to obtain
successively improved guesses for s with the help of a gradient method. After an
integration length of about 100 linear periods P (s = 0 was reached within the
numerical accuracy. Using the transformation (4.14) a solution with zero mean
value was then obtained. The most efficient way to produce a whole set of
solutions was to continue the integration after a small change of S.
For all numerical results illustrated here er = 7 was chosen for the satura-
tion constant. This value corresponds to the adiabatic flame temperature.
Figure 4 shows a set of solutions (p for small inlet losses (2 = 2) and a weak
contraction at the end of the combustion chamber (R = 4). The temperature
266 J. J. Keller, W. Egli and J. Hetlat ZAMP

o . I ~ A 2 A ~'I~ 4
(1

Figure 4
Dimensionless velocity q~ versus di-
mensionless time v for R = 4, 2 = 2
and the values of S listed in Table 1.

00 ~ z/'~ ....~-c 4

(3

Figure 5
Dimensionless velocity ~0 versus di-
mensionless time v for R = 40, 2 = 2
and the values of S listed in Table 1.

ratio S represents the set parameter. Figure 5 illustrates the effect of a strong
exit contraction (which leads to a strong excitation) and Fig. 6 shows the influ-
ence of a large inlet losses. The values (2 = 37, R = 40) chosen for the example
illustrated in Fig. 6 correspond to the measured inlet toss and the geometric exit
contraction of the experimental example which is discussed subsequently.
Vol. 36, 1985 Thermally induced low-frequency oscillations 267

, 2/-'X --,4
t
(2

(4
1
Figure 6
Dimensionless velocity ~0 versus dimen-
(5
sionless time ~ for R = 40, 2 = 37 and the
values of S listed in Table 1.

*o[ _ \ y _\ ,

Figure 7
Solutions (2) and (3) (Figure 6) modified by the inclusion of a time lag, according to (5.10).

In a practical case the response of the temperature is not immediate but


subject to at least a small time lag (or space lag). To investigate the influence
of a time lag A'c on the solutions the upstream boundary condition (4.18),
which represents a relation between particle velocity and soundspeed, was re-
placed by

g(-c + ~,(-c)) = min S + q~(-c Az) (5.10)


aS
whereby the values of A-c was chosen to be 1.5% of the fundamental period.
Figure 7 illustrates the modification due to this time lag of the second and third
solutions shown in Fig. 6. Whilst amplitudes and wavelengths remain essentially
unchanged, the signal shapes are strongly distorted by the inclusion of a time
lag. As a "non-local" character of the upstream boundary condition may also
arise due to a space lag the sign of Az is not necessarily positive. Correspondingly
the distortion of the solutions with respect to the z-axis can be positive or
negative.
268 J. J. Keller, W. Egli and J. Hellat ZAMP

Figures 4 - 7 both show stable and unstable solutions. As was shown in


Sect. 5.1. there is a stable and an unstable branch in the S-d-diagram (i.e.
temperature ratio versus peak-to-peak amplitude). The following table summa-
rizes all relevant data for the numerical solutions and their stability status:

Table 1
Curve Peak-to-peak Temperature Pressure drop Stability
number amplitude d ratio S constant f2 status
Figure 4
(1) 0.100 0.226 4.985 unstable
(2) 0.223 0.228 4.928 stable
(3) 0.222 0.224 4.928 stable
(4) 0.155 0.189 5.079 stable
Figure 5
(1) 0.233 0.324 40.74 unstable
(2) 0.665 0.330 39.94 unstable
(3) 0.668 0.333 39.10 unstable
(4) 0.663 0.338 37.77 stable
(5) 0.621 0.291 40.94 stable
Figure 6
(t) 0.207 0.250 45.31 unstable
(2) 0.288 0.250 45.24 unstable
(3) 0.293 0.250 45.21 stable
(4) 0.294 0.248 45.30 stable
(5) 0.256 0.202 47.95 stable
Figure 7
(2) 0.288 0.250 45.23 unstable
(3) 0.293 0.250 45.21 stable

It is intresting to note that the pressure drop constant f2 seems to have a


minimum at the point which separates the stable from the unstable branch in the
S-d-diagram. Figure 8 shows the peak-to-peak amplitude d versus 1 - S for the
parameter combinations considered. The points corresponding to the solutions
shown in Figs. 4 - 7 are marked by dots.
A comparison of the results, which have been derived analytically in
Sect. 5.1, with the numerical results (compare e.g. Fig. 3 and 8) confirms that
the accuracy of the analytically predicted amplitudes is excellent for large values
of R.
Considering Fig. 8 it may be noted that in the case (R = 40, 2 = 2) the stable
and the unstable solution branch intersect each other and that there is a small
domain where the unstable branch lies above (towards higher values of d) the
stable branch. This is a consquence of having simply chosen the peak-to-peak
amplitude to measure the strength of the oscillation. If we had chosen the energy
of an oscillation to measure its strength (which is the physically appropriate
quantity, but leads to a more complicated numerical formulation of the prob-
lem) the two solution branches would not intersect.
Vol. 36, 1985 Thermally induced low-frequency oscillations 269

1.00
t
"o

LiJ
s

I--
_J 32
s
,<

< 0.50
~J
0_
I
0
p-
I

w
R = 4 , X=2

, ,, ~,,, I ,~I . . . . . "., I ,


0.65 0.75 0.85
1-S
Figure 8
Peak-to-peak amplitude d versus relative temperature increase 1 - S for the parameter combina-
tions (R, 2) = (4, 2), (40, 2), (40, 37). The numbered points on the solution branches refer to the
numerical solutions shown in Figs. 4-7. The stars refer to the experimental data discussed in Sect. 6.

6. Comparsion with experimental results

The apparatus used to produce experimental results which can be compared


to the present theory was a model combustor which had been designed for
qualitative simulation of various effects appearing in an industrial combustion
chamber. Its burner was not particularly well suited to produce the idealized
conditions which have been assumed for the theory. However, the principal aim
of this comparison of experimental and theoretical results was to determine the
qualitative agreement.
The present type of low-frequency oscillations could be observed and identi-
fied in a model combustion chamber (see Fig. 9) of length L = 1100 ram, cham-
ber diameter D = 150 mm, exit diameter D~ = 50 mm. The burner used was a
swirl burner with central propane injection. This arrangement provided a stably
burning diffusion flame with a length always less than 20% of the chamber
length for a wide range of equivalence ratios and Reynolds numbers. Hence the
condition of having a compact zone of combustion was fairly well fulfilled. To
provide a constant rate of fuel addition, the propane supply line contained a
critical-flow nozzle a small distance upstream of the injection nozzle.
The pressure oscillations in the combustion chamber were measured with
water-cooled microphones installed at several axial locations. The air-flow oscil-
lations were measured by means of a hot-wire which was installed upstream of
the swirl burner (position ! in Fig. 9), about one tube diameter downstream of
a large plenum chamber (an order of magnitude larger in volume than the
270 J. J. Keller, W. Egli and J. Hellat ZAMP
upstream
im c h a m b e r

Lance
ane)

body

Figure 9 --
Schematic diagram of the model combustion chamber, dow~....
plenum c h a m b e r

combustion chamber) which provided acoustic isolation of the combustion


chamber on the upstream side. At the location where the velocity was measured
the departures from a top-hat profile were small. Downstream of the exit con-
traction of the combustion chamber there was another large plenum chamber to
acoustically isolate the combustion chamber on the downstream side.
Reaction rate and reaction rate oscillations were determined qualitatively
by measuring the U V-radiation of the OH-radical. Pt-PtRh-thermocouples with
a wire diameter of 25 microns were used to measure temperature oscillations.
The thermocouples were calibrated individually with respect to their transmis-
sion characteristics at every measurement location. For this purpose the volt-
age response of every element to a known electrical current impulse was mea-
sured.
Temperatures and pressures at various axial locations of the combustion
chamber and the air mass flow into the combustor were measured simulta-
neously. Random high-frequency disturbances appeared superimposed on all
low-frequency oscillations. Furthermore, the low-frequency oscillation cycles
showed a certain scatter with respect to period and amplitude. For this reason
the individual periods were first normalized to the average period. Subsequently
an averaged oscillation cycle was computed on the basis of a signal length of
100 periods.
Vol. 36, 1985 Thermally induced low-frequency oscillations 271

Temperature field measurements revealed significant departures from the


quasi-one dimensional flowfield assumed for the analytical investigation. The
zone of flow reversal (downstream of the burner) which is required to stabilize
a compact flame seems to be responsible for phases and amplitudes of the
temperature oscillations showing a degree of non-uniformity over cross-sections
downstream of the reaction zone. The temperature oscillation measurements
were always taken at the radial locations where the largest oscillation ampli-
tudes were observed. These locations were always found outside the vortex
core.
Figure 10 shows the measured phase difference between the fundamental
harmonics of the pressure and temperature oscillations along the combustion
chamber for the fundamental mode. The burner plane and exit contraction are
located at x/L = 0 and x/L = 1, respectively. The experimental results confirm
that the phase difference is zero at the exit contraction and close to ~z at the
burner plane. Furthermore, the phase difference varies linearly along the cham-
ber. Figure 11 shows simultaneously recorded velocity, temperature and pres-
sure signals. Although the velocity was measured at position I (see Fig. 9), it is
scaled to the cross-sectional area of the combustion chamber (see definition of
Uc). The pressure disturbance Ap~ = PM - (,P~t> and temperature Tu were mea-
sured close to the downstream end of the chamber at the axial location marked
by "II" (see Fig. 9), i. e. at x/L = 0.83. In this case the normalized air-to-fuel ratio,
(i. e. the reciprocal of the equivalence ratio), r e = 1.94 and Tc = 295 ~ K, which
approximately corresponds to S = 0.193. For the second example of simulta-
neously recorded signals, shown in Fig. 12, r~ = 2.21 whilst the air temperature
remained unchanged (Tc = 295~ The corresponding temperature ratio
S = 0.214o
From a measurement of static pressures at the wall of the combustion
chamber under cold-flow conditions the loss coefficient Z = 37. As shown in
Fig. 9 the exit contraction r -- A/A N = 9, corresponding to R -- 40. The normal-
ized peak-to-peak velocity amplitudes obtained from a series of measurements,

0.8

0.6
o

:=~ 0 . 4

Figure 10
0.2
Phase delay of the fundamental harmonic of the
negative pressure disturbance with respect to the
fundamental harmonic of the temperature distur- 0
bance in fractions of an oscillation period versus 0 0.2 0.4 0.6 0.8
dimensionless axial location x/L. • / L
272 J. J. Keller, W. Egli and J. Hellat ZAMP

[m/s]|
3.84 . . . . . . . . . . . . . . . . . . .

u c
0.1 0.2 O.3~t

t ,,
2.04 ~ . . . . ~ . . . . . . ~ . . . . . . ~ . . . .

[ o K ]
t436 . . . . . . . . . . . . . . . . . . . .
TH
t t249 I I
1095 . . . .

Figure 11 [m bor]/
Velocity (scaled to the cross-sectional area of 5.8 . . . . . . . . . . . . . . . . . .
the combustion chamber) measured at posi-
tion I (see Fig. 9), temperature and pressure
disturbance measured at position II at the nor-
t -6. m . . . . . .
malized air-to-fuel ratio r~ = 1.94 (S = 0.193).

[m/s]
4.70.t.~. . . . . . . . . . . . .A. .-. -. . . . -~-
uC
0,t 0.2 0.3 ~ t
t
2.04 . . . . . . .

[oK]~ _
t597 . . . . . . . . . . . . . . . . . . .

TH

f 1083

858- ---

igure2
Velocity (scaled to the cross-sectional area of
mb~ 98-
the combustion chamber) measured at posi- ~PM
tion I (see Fig. 9), temperature and pressure t o I
disturbance measured at position II at the nor-
malized air'to-fuel ratio r E = 2.21 (S = 0.214). -12.4-

including the cases illustrated in Figs. 11 and 12, are marked by small stars in
Fig. 8. The flow in the combustion chamber is not free of swirl. As a consequence
we cannot expect a very good agreement with the amplitude curve (R = 40,
2 = 37). However, as should be the case, the orders of magnitude of the ampli-
tudes do indeed agree, the temperature domain where this low-frequency insta-
bility appears is rather accurately predicted by the analysis and the amplitude
decrease due to temperature saturation with decreasing values of S is also
Vol. 36, 1985 Thermally induced low-frequency oscillations 273

observed. Furthermore, comparing pressure disturbance and temperature sig-


nals in Figs. 11 and 12 it can be seen that the downstream reflection condition
is reasonably well described by (3.3). Although the effect of temperature satura-
tion in reality is clearly not as abrupt as defined by the upstream boundary
condition (4.18), the pressure disturbance and velocity signals in Fig. 11 exhibit
saturation features.

7. Concluding remarks

As implied by the title of the paper, the appearance of the present kind of
low-frequency oscillation requires heat addition to the flow system considered.
Although combustion is one of several possibilities to provide a sufficiently
large rate of heat addition, it probably represents the most interesting applica-
tion.
A further point to be discussed concerns the extension of the present ideas
to the more general case of axially distributed heat addition (distributed com-
bustion). A correspondingly generalized linear stability analysis shows two ma-
jor differences with respect to the special case of concentrated combustion. As
might be expected the fundamental mode is not necessarily the most unstable in
cases of distributed heat addition. Depending upon the properties of the distri-
bution of heat addition certain higher modes may be preferred. Furthermore, the
response of the entropy modulation to a pressure disturbance is never stronger
than in the case of concentrated combustion. A departure from the presently
assumed idealisation, which is more relevant in practice, is that of distributed air
injection (e. g. film cooling).
If secondary air is injected at an axial location where the temperature and
pressure oscillate in phase, the entropy wave is amplified by the secondary air
injection and the instability may be enhanced.

References

[1] L. Crocco and S. I. Cheng, Theory of combustion instability in liquid propellant rocket motors,
AGARDograph No. 8, Butterworths 1956.
[2] C. E. Mitchell, L. Crocco and W. A. Sirignano, Nonlinear longitudinal instability in rocket mo-
tors with concentrated combustion. Combust. Sci. Technol. 1, 35 (1969).
[3] S. M. Candel, Analyticalstudies of some acoustic problems ofjet engines. Ph.D. thesis, California
Institute of Technology, Pasadena 1972.
[4] F. E. Marble, Acoustic disturbance from gas nonuniformities through a nozzle. Proc. Interagency
Symp. Univ. Res. Transp. Noise, Stanford University, June 1973, Vol. 3 (1973).
[5] N. A. Cumpsty and F. E. Marble, The interaction ofentropyfluctuations with turbine blade rows;
a mechanism of turbojet engine noise. Proc. R. Soc. Lond. A. 357, 323 (1977).
[6] J. E. Ffowcs Williams and M. S. Howe, The generation of sound by density inhomogeneities in low
Mach number nozzle flows. J. Fluid Mech. 70, 605 (1975).
274 J. J. Keller, W. Egli and J. Hellat ZAMP

[7] F. E. Marble and S. M. Candel, Acoustic disturbance from gas nonuniformities convected through
a nozzle. J. Sound Vibr. 55, 225 (1977).
[8] J. J. Keller, Production and propagation of sound in a duct. J. Acoust. Soc. Am. 65, 25 (1979).
[9] J. J. Keller, Third order resonances in closed tubes. Z. Angew. Math. 27, 303 (1976).

Summary
This paper discusses nonlinear thermally induced oscillations in quasi-one dimensional flows
in ducts. In practice such oscillations are frequently observed in furnaces and combustion chambers.
The problem considered involves entropy disturbances which are convected through a nozzle at the
end of a tube, thereby producing an acoustic wave which propagates upstream and leads to a
modulation of the mass flow at the inlet of the tube. Alternatively, if the rate of heat addition (i. e.
the rate of fuel addition in the case of a combustion chamber) responds only weakly or not at all
to the oscillating pressure at the inlet, the modulated air flow produces an entropy oscillation (due
to the oscillating equivalence ratio in the case of a combustor) downstream of the zone of heat
addition (reaction zone). To obtain general stability limits for this kind of self-induced oscillation,
a second-order analysis is developed which leads to a nonlinear wave equation. The convection of
entropy disturbances introduces nonlinear memory effects which are responsible for a non-local
character of the wave equation. The wave equation is solved with the help of a numerical evolution
scheme, making use of a suitable scaling transformation which does not change the form of the
equation.

Zusammenfassung
Es werden nichtlineare thermisch getriebene Schwingungen in quasieindimensionalen Rohren
untersucht. In der Praxis kann dieser Schwingungstyp h/iufig in Feuerungsanlagen und Brennkam-
mern beobachtet werden. Die betrachteten Schwingungen werden yon Entropiest6rungen angetrie-
ben, welche die Austrittsdiise eines Rohres durchqueren und dabei Schall erzeugen. Eine der
erzeugten Schallwellen l/iuft stromaufw/irts und verursacht eine Massenstrommodulation am Ein-
tritt des Rohres. Falls die W/irmezufuhrrate (Brennstoffzufuhrrate in einer Brennkammer) nur
schwach oder gar nicht auf Druckschwankungen am Eintritt reagiert, dann verursacht die Luft-
massenstromschwankung eine Entropieschwankung (aufgrund des schwankenden Brennstoff/Luft-
Verhgltnisses) am Ende der Zone (Reaktionszone), in welcher W/irme zugefiihrt wird. U m allge-
meine Stabilitfitsgrenzen f/ir solche selbsterregten Schwingungen zu ermitteln, wird eine Theorie
zweiter Ordnung entwickelt, die zu einer nichtlinearen Wellengleichung fiihrt. Die Konvektion von
Entropiest6rungen f/ihrt zu nichtlinearen Ged/iehtniseffekten, die fiir den nicht-lokalen Charakter
der Wellengleichung verantworlich sind. Die Gleichung wird mit Hilfe eines numerischen Evolutions-
schemas gel6st, wobei von einer Skalierungstransformation Gebrauch gemacht wird, welche die
Form der Gleichung nicht ver/indert.

(Received: March 6, 1984)

You might also like