linear thermoacoustic instability in the time domain

9
Linear thermoacoustic instability in the time domain S. Karpov and A. Prosperetti Department of Mechanical Engineering, The Johns Hopkins University, Baltimore, Maryland 21218 ~Received 30 April 1997; accepted for publication 17 February 1998! An approximate time-domain description of the development of the thermoacoustic instability in gas-filled tubes is developed by exploiting the difference between the instability time scale and the period of standing waves. The perturbation results compare very favorably with the exact frequency-domain theory of Rott. The perturbation results are further simplified by introducing a short-stack approximation which is numerically much simpler and only slightly less accurate. An approximate expression for the critical temperature gradient accounting for viscous effects and other design features is also derived. In addition to the fundamental mode of a tube closed at both ends, the theory includes higher modes as well as open-end boundary conditions. © 1998 Acoustical Society of America. @S0001-4966~98!00606-7# PACS numbers: 43.35.Ud @HEB# INTRODUCTION The linear theory of thermoacoustic effects as developed in the well-known series of papers by Rott 1–4 ~good reviews are provided by Wheatley 5 and Swift 6 ! is based on a formu- lation in the frequency domain. In the present paper we de- velop an approximate time-domain approach that is valid when the period of standing waves is much shorter than the growth rate of the instability. Since this condition is com- monly met in practice, we expect the present formulation to furnish a realistic description of the time evolution of the instability in a wide variety of situations. One of the main results of the present work is an expres- sion for the linear growth rate of the thermoacoustic instabil- ity or, below threshold, for the decay rate of standing waves. This latter result can also be viewed as expressing the power requirement for the operation of a thermoacoustic refrigera- tor. In the past these results have been approached by esti- mating the so-called work flux ~see, e.g., Ref. 6! or Q -value of the device. 7–9 Our method may therefore be interpreted as a different approach to the evaluation of these quantities. Its advantage is a greater flexibility that enables us to consider- ably simplify their quantitative evaluation with respect to the exact theory of Rott while at the same time maintaining a greater accuracy than existing approximations. A second result of the paper is an unambiguous defini- tion of the concept of critical temperature gradient when vis- cous and thermal effects are both important and an accurate approximate expression for this quantity. I. MATHEMATICAL MODEL In Watanabe et al. 10 it was shown that a model of ther- moacoustic devices that, in the frequency domain, is exactly equivalent to Rott’s, may be written in the form ]r ] t 1 1 S ] ] x ~ S r 0 u ! 50, ~1! r 0 ] u ] t 1 ] p ] x 52r 0 D ~ u ! , ~2! ] p ] t 1 g p 0 S ] ~ Su ! ] x 5~ g 21 ! F H~ T w 2T ! 2 dT w dx Q ~ u ! G . ~3! The system of equations is closed by assuming the validity of the perfect gas law. Here r, p , T , and u are the density, pressure, temperature, and velocity fields. The spatial coor- dinate x is directed along the axis of the device, S ( x ) is the ~possibly variable! cross-sectional area, and g the ratio of specific heats. The subscript zero denotes undisturbed quan- tities. The temperature T w ( x ) of the solid surfaces in contact with the gas ~i.e., the tube walls and the stack plates! is assumed to be only a function of x and is taken to be pre- scribed in the following. It should be noted that, even when this quantity evolves with time, it does so only slowly ~com- pared with the period of the oscillations! and therefore the present method can be extended to deal with this case as well as noted below. The drag operator D and the heat transfer operators H and Q are given, in the frequency domain, by D ˜ ~ u ! 5i v f V 1 2 f V u ˜ , ~4! H ˜ ~ T w 2T ! 52i vr 0 c p f K 1 2 f K T ˜ 8 , ~5! Q ˜ ~ u ! 5r 0 c p F 1 1 2s S 1 1 2 f V 2 s 1 2 f K D 21 G u ˜ . ~6! Here T 8 5T 2T w and tildes denote frequency-domain quan- tities. The parameters f V, K depend on the ratio of the diffu- sion lengths d V, K to the plate spacing l and, for disturbances proportional to exp ivt ˆ , are given by 6 f 5~ 1 2i ! d l tanh~ 1 1i ! l 2 d , ~7! where the index can be V or K . The viscous and thermal penetration lengths d V, K are given by 3309 3309 J. Acoust. Soc. Am. 103 (6), June 1998 0001-4966/98/103(6)/3309/9/$10.00 © 1998 Acoustical Society of America Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Upload: a

Post on 15-Apr-2017

219 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Linear thermoacoustic instability in the time domain

Redistrib

Linear thermoacoustic instability in the time domainS. Karpov and A. ProsperettiDepartment of Mechanical Engineering, The Johns Hopkins University, Baltimore, Maryland 21218

~Received 30 April 1997; accepted for publication 17 February 1998!

An approximate time-domain description of the development of the thermoacoustic instability ingas-filled tubes is developed by exploiting the difference between the instability time scale and theperiod of standing waves. The perturbation results compare very favorably with the exactfrequency-domain theory of Rott. The perturbation results are further simplified by introducing ashort-stack approximation which is numerically much simpler and only slightly less accurate. Anapproximate expression for the critical temperature gradient accounting for viscous effects and otherdesign features is also derived. In addition to the fundamental mode of a tube closed at both ends,the theory includes higher modes as well as open-end boundary conditions. ©1998 AcousticalSociety of America.@S0001-4966~98!00606-7#

PACS numbers: 43.35.Ud@HEB#

pe

-dlith-te

ebiewrae

a. Idehe

nisra

-ct

dity

oor-

uan-t

-en

well

n--

l

INTRODUCTION

The linear theory of thermoacoustic effects as develoin the well-known series of papers by Rott1–4 ~good reviewsare provided by Wheatley5 and Swift6! is based on a formulation in the frequency domain. In the present paper wevelop an approximate time-domain approach that is vawhen the period of standing waves is much shorter thangrowth rate of the instability. Since this condition is commonly met in practice, we expect the present formulationfurnish a realistic description of the time evolution of thinstability in a wide variety of situations.

One of the main results of the present work is an exprsion for the linear growth rate of the thermoacoustic instaity or, below threshold, for the decay rate of standing wavThis latter result can also be viewed as expressing the porequirement for the operation of a thermoacoustic refrigetor. In the past these results have been approached bymating the so-called work flux~see, e.g., Ref. 6! or Q-valueof the device.7–9 Our method may therefore be interpreteda different approach to the evaluation of these quantitiesadvantage is a greater flexibility that enables us to consiably simplify their quantitative evaluation with respect to texact theory of Rott while at the same time maintaininggreater accuracy than existing approximations.

A second result of the paper is an unambiguous defition of the concept of critical temperature gradient when vcous and thermal effects are both important and an accuapproximate expression for this quantity.

I. MATHEMATICAL MODEL

In Watanabeet al.10 it was shown that a model of thermoacoustic devices that, in the frequency domain, is exaequivalent to Rott’s, may be written in the form

]r

]t1

1

S

]

]x~Sr0u!50, ~1!

r0

]u

]t1

]p

]x52r0D~u!, ~2!

3309 J. Acoust. Soc. Am. 103 (6), June 1998 0001-4966/98/103

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

d

e-de

o

s-l-s.er-

sti-

stsr-

a

i--te

ly

]p

]t1

gp0

S

]~Su!

]x5~g21!FH~Tw2T!2

dTw

dxQ~u!G .

~3!

The system of equations is closed by assuming the valiof the perfect gas law. Herer, p, T, andu are the density,pressure, temperature, and velocity fields. The spatial cdinatex is directed along the axis of the device,S(x) is the~possibly variable! cross-sectional area, andg the ratio ofspecific heats. The subscript zero denotes undisturbed qtities. The temperatureTw(x) of the solid surfaces in contacwith the gas~i.e., the tube walls and the stack plates! isassumed to be only a function ofx and is taken to be prescribed in the following. It should be noted that, even whthis quantity evolves with time, it does so only slowly~com-pared with the period of the oscillations! and therefore thepresent method can be extended to deal with this case asas noted below.

The drag operatorD and the heat transfer operatorsH

andQ are given, in the frequency domain, by

D~u!5 ivf V

12 f Vu, ~4!

H~Tw2T!52 ivr0cp

f K

12 f KT8, ~5!

Q~u!5r0cp F 1

12s S 1

12 f V2

s

12 f KD21G u. ~6!

HereT85T2Tw and tildes denote frequency-domain quatities. The parametersf V,K depend on the ratio of the diffusion lengthsdV,K to the plate spacingl and, for disturbancesproportional to expivt, are given by6

f 5~12 i !d

ltanh~11 i !

l

2d, ~7!

where the index can beV or K. The viscous and thermapenetration lengthsdV,K are given by

3309(6)/3309/9/$10.00 © 1998 Acoustical Society of America

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 2: Linear thermoacoustic instability in the time domain

n

-

m

n

ra

cut

tei

eosthllahe

hesemtha-nduaie

r-

.alin

us

wee

e

ree

en-

Redistrib

dV5A2n

v, dK5A2a

v5

dV

As, ~8!

with n the kinematic viscosity,s the Prandtl number, anda5n/s the thermal diffusivity. For a circular cross sectiowith radiusr 0 one has instead1

f 52J1„~ i 21!~r 0 /d!…

~ i 21!~r 0 /d!J0„~ i 21!~r 0 /d!…, ~9!

again valid for bothf V and f K . When the diffusion penetration depths are small compared with eitherl or r 0 , f givenby either~7! or ~9! admits the asymptotic approximation

f .~12 i !d

l, f .~12 i !

d

r 0. ~10!

More generally, for other cross section shapes, in this lione may setf .2(12 i )d/dh , wheredh54S(x)/s(x) is thehydraulic diameter defined in terms of the cross-sectioareaS and the ‘‘wetted’’ perimeters.

If Eqs. ~1!–~3! are written in the frequency domain, aftereduction to a single equation by differentiation and elimintion, one recovers Rott’s equation:

1

S

d

dx F ~12 f V!Sdp

dxG11

Tw

dTw

dx S 11s f V2 f K

12s Ddp

dx

1v2

c2 @11~g21! f K# p50. ~11!

The local adiabatic speed of soundc(x) equalsAgRTw(x),with R the universal gas constant divided by the gas molelar mass, and is therefore also variable in general alongaxis of the system.

II. THE STABILITY CALCULATION

Experiment~see, e.g., Refs. 5, 8! shows that the initialbuild-up of the thermoacoustic instability has the characof a modulated standing wave the frequency of whichessentially dictated by the resonator, while the amplitudslowly varying in time. This observations suggests the psibility of setting up a perturbation scheme based onsmallness of the ratio of the characteristic period of oscition to the characteristic time for the development of tinstability.

In the framework of the previous model, the terms in tleft-hand side of Eqs.~1!–~3! describe linear acoustic wavein a gas column with variable cross-sectional area and tperature stratification and are therefore responsible for‘‘carrier’’ frequency of the wave. The heart of the thermocoustic effect is in the terms in the right-hand sides amore specifically, in the right-hand side of the energy eqtion ~3!. The observed slowness of the modulation implthat these terms are small or, sinceD , H, andQ all vanishwith f V,K , that thef ’s are small. In order to set up a pertubation scheme, we take thereforef V,K to be of ordere!1and we will treat the ratiof V,K /e as a quantity of order 1While a formal definition ofe is not necessary as the finresults do not explicitly depend on this parameter, accordto ~10!, one may think of it as the ratiod/ l .

3310 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

it

al

-

-he

rsis-e-

-e

,-

s

g

To proceed with a formal development of the previoobservation, we introduce the scaled times

t5t, t5et, ~12!

and, according to the method of multiple time scales,treat t andt as independent variables. The original variabltis recovered at the end of the calculation by using~12! ~see,e.g., Refs. 11, 12!. As a consequence of these definitions whave, correct to ordere,

]

]t5

]

] t1e

]

]t. ~13!

We also expand the dependent variablesp, u, etc., in apower series ine; for example,

p5p11ep21¯ , ~14!

with a similar notation foru, etc. Now these expansions asubstituted into~1! to ~3! where~13! is used to express thtime derivative. Upon separating orders, to zero order ine,we have

]r1

] t1

1

S

]

]x~Sr0u1!50, ~15!

r0

]u1

] t1

]p1

]x50, ~16!

]p1

] t1

gp0

S

]~Su1!

]x50, ~17!

and, at the next order,

]r2

] t1

1

S

]

]x~Sr0u2!52

]r1

]t, ~18!

r0

]u2

] t1

]p2

]x52

]u1

]t2

1

er0D~u1!, ~19!

]p2

] t1

gp0

S

]~Su2!

]x52

]p1

]t2~g21!

3F1

eH~T18!1

1

eQ~u1!

dTw

dxG .

~20!

A. Zero order

The three equations~15! to ~17! can be combined to give

1

S

]

]xS c2S

]p1

]xD 2

]2p1

] t250. ~21!

The solution of this equation may be taken in the form

p1~x, t,t!5A~t!P1~x!exp~ iv t !1c.c., ~22!

where c.c. denotes the complex conjugate, and the eigfunction P1(x) and eigenfrequencyv are determined bysolving the eigenvalue problem

3310S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 3: Linear thermoacoustic instability in the time domain

wth

ar,al

t

n

n

pe

nt

a-s

go-,

f-

ex-

re, we

re

-

ct

Redistrib

1

S

d

dx S c2SdP1

dx D1v2P150, ~23!

subject to appropriate boundary conditions. In this paperonly consider rigid or open terminations at the ends oftube and therefore

dP1

dx50 or P150 at x50, x5L. ~24!

With these conditions, it can readily be shown by standtechniques that the eigenvaluev is real ~see, e.g., Refs. 1415!. The eigenfunctionP1 can therefore also be taken reand, for later convenience, we normalize it so that

E0

L

S~x!P12~x!dx5Vp0

2, ~25!

where V is the volume of the device andp0 is the undis-turbed static pressure. Since the total acoustic energy indevice, to the present approximation, is given by~see, e.g.,Ref. 16 section 63!

E5E0

L

S~x!^p1

2&r0c2 dx5

A2

2gp0E

0

L

S~x!P12dx5A2E0 ,

~26!

we see that normalization~25! is equivalent to normalizingthe energy by

E05Vp0

2g. ~27!

It is readily found thatu1 , r1 , andT18 also satisfy Eq.~21! and therefore they can also be written in the form~22!with the samev and proportionality to the same functioA(t). In particular, from~16!,

U1~x!5i

vr0

dP1

dx. ~28!

B. First order

To ordere, the drag and heat transfer operatorsD , H,andQ operate on the variablet and, since the dependence othis variable is only through the exponential factor expivt,one can use the Fourier-space representation of these otors given in~4!–~6!. Upon substitution of solution~22! forp1 and similar expressions for the other first-order fields i~18!–~20!, we thus find

1

S

]

]xS c2S

]p2

]xD 2

]2p2

] t2

5H 2ivdA

dtP12

A

e F1

S

]

]x S i f V

12 f Vvr0Sc2U1D

1~g21!v2f KP1~x!

2 if V2 f K

~12s!~12 f V!g

vp0U1

Tw

dTw

dx G J exp iv t1c.c.

~29!

3311 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

ee

d

he

ra-

o

The forcing at frequencies6v in the right-hand side willgenerate resonant terms proportional tot exp ivt in the so-lution for p2 which lead to a breakdown of the approximtion over times of order (ev)21. To avoid this resonance, ain the standard procedure,11–13 we impose the solvabilitycondition that the right-hand side of the equation be orthonal to the solution of the~adjoint! homogeneous equationnamely exp(2ivt )P1:

E0

L

S@exp~2 iv t !P1#S 2ivdA

dtP12

A

e$¯% Dexp~ iv t !dx

50, ~30!

where, for brevity, we write$¯% to denote the coefficient oA in ~29!. Multiplication by S(x) before integration is necessary so that thex-operator in the left hand side of~29! beself-adjoint with the boundary conditions~24!. In this waywe find

dA

dt1

V

eA50, ~31!

where

V5 12~MD1 iND1MH1 iNH! ~32!

with, after an integration by parts,

MD1 iND5i

Vp02v

E0

L

Sc2f V

12 f VS dP1

dx D 2

dx, ~33!

MH1 iNH5 ig21

Vp02v

E0

L

dx SP1F f Kv2P1

2f K2 f V

~12s!~12 f V!cp

dTw

dx

dP1

dx G , ~34!

wherecp is the specific heat at constant pressure.Since thef ’s are essentially the small parametere, and

since these results have an accuracy of first order ine, onemight be tempted to replace 12 f V simply by 1 in the de-nominators. While justified in principle in the limite→0, wehave found that the numerical accuracy of these resultstends to significantly larger values ofd/ l if the forms givenabove are retained.

Clearly ~34! is also applicable whenTw depends slowly~i.e., on the time scalet! on time. Integration of the ampli-tude equation~31! in this case is however somewhat mocomplicated and therefore, as already mentioned beforelimit ourselves to the simpler situation in whichTw is inde-pendent of time. In this case the solution of~31! is

A~t!5A0 expS 2Vt

e D5A0 exp~2Vt !, ~35!

with A0 dependent on the initial conditions, and therefogrows or decays according to the sign ofMD1MH . Thesecond form of solution~35! demonstrates explicitly the independence of the result from the definition ofe. The previ-ous results imply the following approximation for the exaeigenfrequencyv in Rott’s equation~11!:

v.v1 iV. ~36!

3311S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 4: Linear thermoacoustic instability in the time domain

dn-o

,

anethrttthh

e

ti

se

as

by

gthso

con-

eur-gingifi-

om

onsenceeing.

Redistrib

It is interesting to verify that~35! reduces to the standarexpressions for the viscous and thermal damping of a staing wave in a hollow isothermal cylindrical tube. If we consider the case of rigid terminations at both ends, by the nmalization~25!, we may write

P15&p0 cosnpx

L. ~37!

With this expression, noting thatv5npc/L, and using ap-proximation~10! for f V,K , it is immediate to verify that, tolowest order in these quantities,

MD1 iND.~11 i !A2nv

r 0, ~38!

MH1 iNH.~11 i !~g21!A2av

r 0, ~39!

in agreement with well-known results~see, e.g., Ref. 17p. 534!.

In view of the proportionality ofE to theA2 shown by~26!, we also see that~31! implies

dE

dt52~MD1MH!E . ~40!

In the unstable caseMD1MH,0 and this relation gives thegrowth rate of the disturbance energy. In the stable cMD1MH.0 and, in order to maintain the oscillations, oneeds to supply energy externally at the rate dictated byright hand side of~40!. In this case the energy lost is in padissipated by viscosity and thermal conduction, but in paralso the work needed to transport heat from the colder towarmer regions as, in this case, the device behaves as apump, or acoustic refrigerator.

Swift6 defines the work fluxW2 in a thermoacousticengine. This is equivalent to the rate of growth~or decay! ofthe total acoustic energy in the device and, in the presnotation, is therefore given by

W252~MD1MH!E . ~41!

The connection between Swift’s expression forW2 and ourresult is best explained after the considerations of SubsecIII A.

The quality factorQ is defined by

Q52vE

W2

, ~42!

which, with ~32! and ~41!, becomes

Q5v

2 ReV. ~43!

This relation establishes a connection between the prestudy and the work of Atchley7–9 who estimated the growthrate of the instability using an approximate calculation ofE

andW2 . The present approach has the advantage of gregenerality which, as will be seen in the following sectionleads to very accurate results.

3312 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

d-

r-

se

e

iseeat

nt

on

nt

ter,

III. SHORT-STACK APPROXIMATION

The asymptotic results of Sec. I can be approximatedadopting the so-called ‘‘short-stack approximation’’~see,e.g., Ref. 6!. We thus assume that the stack occupies a lenmuch smaller than the wavelength of the standing wavesthat the pressure field in the stack can be considered asstant. We takef V,K.0 outside the stack region.

We thus write~33!, approximately, as

MD1 iND.i f V

12 f V

VS

V

c2

vp02 S dP1S

dxD 2

. ~44!

Here f V and c2 ~or, equivalentlyTw! are evaluated at themean temperature in the stack,Tw5 1

2(TC1TH). It is diffi-cult on the basis ofa priori considerations to formulate thkind of averaging that is most effective for the present pposes. We have tried several definitions, such as averacomputed by integration along the stack, but found insigncant differences. Furthermore,

VS5ExS2~1/2!LS

xS1~1/2!LSdx S~x! ~45!

is the volume of the gas in the stack region extending frxS2 1

2LS to xS1 12LS . The indexS appended toP1 denotes

evaluation at the stack’s midpointxS . In a similar way, from~34!,

MH1 iNH. iVS

V

v

p02 F ~g21! f K~P1S!2

2f K2 f V

~12s!~12 f V!

1

v2

dc2

dxP1S

dP1S

dxG ,

~46!

and, upon combining results~44! and ~46!,

MD1MH.VS

V

v

p02 FZ1

c2

v2 S dP1S

dx D 2

1~g21!Z2~P1S!2

2Z3

1

v2

dc2

dxP1S

dP1S

dx G , ~47!

where we have set for brevity

Z152ImS f V

12 f VD , Z252Im~ f K!,

~48!

Z351

12sImS f V2 f K

12 f VD .

For small diffusion penetration depths, from~10!, thesequantities are given by

Z1.dV

l, Z2.

1

As

dV

l, Z3.

1

As~11As!

dV

l. ~49!

Use of these approximations in actual numerical calculatiis not recommended. We show them to stress the dependof the quantitiesZj on the Prandtl number and the ratio of thviscous boundary layer thicknesses to the the plate spac

3312S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 5: Linear thermoacoustic instability in the time domain

ycxou

y:

aoa

the

ra

at

‘e

inof

ale.the

ceipa-is

be

is

ra-

t

isedical

Redistrib

From definitions~41! of W2 and~22! of p1 , and relation~26!we have

W252VSv

gp0FZ1

c2

v2 K S ]p1S

]x D 2L 1~g21!Z2

3^~p1S!2&2Z3

1

v2

dc2

dx K p1S

]p1S

]x L G , ~50!

where the angle brackets denote time average over one cIn his Eq.~80! Swift6 gives an expression for the work fluW2 . It can be shown that his expression is the same asprovided that f V,K in ~48! are approximated by~10! anddTw /dx is much larger thanTw /L. Since usuallyTH2TC isnot very much smaller thanTw , while LS!L, this latterapproximation is often satisfied.

By equating~47! to zero we find a critical value of thetemperature gradient that, if exceeded, leads to instabilit

dTw

dx Ucrit

51

Z3FZ2

v2

cp

P1S

dP1S /dx1Z1Tw

dP1S /dx

P1SG . ~51!

If viscous effects are neglected,f V50, s50 and this expres-sion is readily shown to coincide with the critical temperture gradient defined in the elementary theory of thermcoustic processes~see, e.g., Ref. 6!. The minimum ofdTw /dxucrit occurs for

P1S

dP1S /dx5

1

v FZ1

Z2cpTwG1/2

, ~52!

where it has the value

dTw

dx Umin

52vAZ1Z2

Z3S Tw

cpD 1/2

. ~53!

In particular, in the inviscid case, any nonzero value oftemperature gradient will give rise to an instability if thstack is positioned at a pressure node.

To obtain further explicit results, we distinguish sevetypes of boundary conditions.

A. Closed tube

Use of the previous formulae requires the approximknowledge ofv andP1 in the stack region. We calculatevfrom

v2.S znp

L D 2

gRTe , ~54!

where Te is an effective temperature andz is a factor ac-counting for the difference between the actual and the ‘fective’’ length of the tube~see, e.g., Ref. 18 art. 265!. Wehave used Rayleigh’s method to estimatez, but found a neg-ligible difference and therefore we set this quantity to 1the following. We takeTe to be the average temperaturethe system

Te51

L E0

L

Tw~x!dx. ~55!

We approximate thenth pressure eigenmodeP1 in the stackregion by

3313 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

le.

rs

--

e

l

e

f-

P15&p0 cosv

cH~L2x!, ~56!

where cH5AgRTH and the factor& is suggested by theform of ~37!. Taking the wave number asv/cH rather thannp/L is motivated by the result shown later that the optimstack position for instability is near the hot end of the tubWe find that this choice indeed improves agreement withexact solution as expected.

With these approximations we find from~47!

MD1MH.2VS

VvH Z1

Tw

THsin2F v

cH~L2xS!G

1~g21!Z2 cos2F v

cH~L2xS!G

21

2Z3

cH

vTH

dTw

dxsinF2v

cH~L2xS!G J . ~57!

The first two terms are positive definite—and henstabilizing—and correspond to viscous and thermal disstion, respectively. The only potentially destabilizing termthe last one. If, as we assume,dTw /dx>0, it is thereforeevident that, for instability to be possible, the stack mustpositioned in a region where sin 2(v/cH)(L2xS).0. In par-ticular, for the fundamental mode for whichn51, the stackmust be placed to the right of the tube’s midpoint. In thcase it is easy to see thatTw.Te .

With expression~54! for v, the smallest critical tem-perature gradient~53! becomes

dTw

dx Ucrit

52npTe

L

AZ1Z2

Z3A~g21!

Tw

Te. ~58!

This relation shows that, for a given mean temperature gdient, only a finite number of modes can be unstable.

Expression~57! can be identically rewritten as

MD1MH.VS

VvZ2FZ1Tw

Z2TH1g21

1B cosS 2v

cH~L2xS!1f D G , ~59!

where

B5F S g212Z1Tw

Z2THD 2

1S Z3cH

Z2vTH

dTw

dx D 2G1/2

, ~60!

sin f51

B

Z3

Z2

cH

vTH

dTw

dx. ~61!

Upon ignoring the small effect of the dependence ofv onxS , it is evident from~59! that the instability will be greatesat the position where the cosine equals21, from which

sinF2v

cH~L2xS!G5sin f. ~62!

It is readily verified that, if the mean pressure gradientvery small, this relation requires that the stack be positionnear a pressure node, as found before. However, in typ

3313S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 6: Linear thermoacoustic instability in the time domain

acmth

e

c

,

ityrtott

editreious

.u

seon.

-

ofube

een

mane

eenn theale is

Redistrib

conditions of operation of thermoacoustic devices, the stlength is much smaller than the tube length, while the teperature difference along the stack is of the order ofcold-end absolute temperature.6 As a consequence,dTw /dxis usually much larger thanTe /L so that, for moderate modnumbern, we may approximateB andf by

B.Z3

Z2

cH

vTH

dTw

dx, sin f.1. ~63!

With these approximations, the optimal position of the stato destabilize thenth mode is very nearly sin 2(v/cH)(L2xS)51, i.e.,xS /L512(k11/4)cH /(nce), with k an inte-ger subject to the only restrictions that 0<xS /L<1. In par-ticular, for the fundamental moden51, k must be taken as 0so that

xS

L512

1

4ATH

Te. ~64!

In the elementary theory of the thermoacoustic instabilthe optimal stack position is found to be at the three-quapoint xS5 3

4L, i.e., one-eighth of a wavelength from the hend~Ref. 6; recall that here distances are measured fromcold-end of the tube!. SinceTH.Te , our result~64! showsthat the optimal stack position is actually slightly displacfurther away from the hot end of the tube, in agreement wthe exact results shown in Figs. 1 and 2. It should bemarked that this conclusion only applies under the conditthat udTw /dxu@Te /L and therefore it can lead to erroneoconclusions whenudTw /dxu is small.

As the ordern of the mode increases,v also increasesand the constantB given by ~63! accordingly decreasesThus higher-order modes are found to be always lessstable than the fundamental one.

FIG. 1. Imaginary part of the eigenfrequency Imv as a function of the stackpositionxS /L along the resonant tube. The temperature difference betwthe two ends of the stack is 368 K and the gas pressure 307 kPa. Whestack is positioned at the midpoint of the tube the ratio of the therpenetration length to the spacing of the stack plates is 0.34. The solid lithe exact result from~11!, the dotted line the two-time-scales result~32!, thedashed line the short-stack approximation~57!, and the dash-dot line Swift’sresult. Instability corresponds to Imv,0.

3314 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

k-e

k

,er

he

h-n

n-

B. Tube open at one end

For a tube open at the cold endx50 and rigidly termi-nated atx5L, we approximate thenth eigenfrequency by

v2.F S n11

2D p

L G2

gRTe , ~65!

where the average temperatureTe is defined as before by~55!. The corresponding eigenmodeP1 is taken as

P15&p0 sinv

cCx, ~66!

wherecC5AgRTC. Unlike the previous case, here we uv/cC for the wave number since the optimal stack positifor instability will be found to be close to the cold endProceeding as before, we find

MD1MH.2VS

VvFZ1

Tw

TCsin2 S v

cCxSD

1~g21!Z2 cos2 S v

cCxSD

21

2Z3

cC

vTC

dTw

dxsin2 S v

cCxSD G . ~67!

With ~65!, the critical gradient~53! is

dTw

dx Ucrit

52S n11

2D pTe

L

AZ1Z2

Z3A~g21!

Tw

Te. ~68!

When udTw /dxu@Te /L the maximum instability occurs approximately atxS /L5 1

2@(4k11)/(2n11)#(cC /ce), wherece5AgRTe. For the lowest moden50, k50 and xS

5 12L(cC /ce). This result is easily understood on the basis

the previous one as the present situation is similar to a tof length 2L closed at both ends.

nthelis

FIG. 2. Imaginary part of the eigenfrequency Imv as a function of the stackpositionxS /L along the resonant tube. The temperature difference betwthe two ends of the stack is 368 K and the gas pressure 307 kPa. Whestack is positioned at the midpoint of the tube the ratio of the thermpenetration length to the spacing of the stack plates is 0.17. The solid linthe exact result from~11!, the dotted line the two-time-scales result~32!, thedashed line the short-stack approximation~57!, and the dash-dot line Swift’sresult. Instability corresponds to Imv,0.

3314S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 7: Linear thermoacoustic instability in the time domain

t t

tiv

eth

a

.

-

leob

nts

cmthethw

tthans

or

w

rhnteo

el

f

r

av-gth,es

unc-

d-ken

isre-ow

surethethe

05heell

atioe

.34theepthig.

he

e

ive.

rgeatfor

theof

gapft’s

inion

re

-vi-

r in

Redistrib

The case of a tube closed at the cold end and open aother one is contained in the preceding formulae ifdTw /dxis taken to be negative. The critical gradient is the negaof ~68!. Again with the hypothesis thatudTw /dxu@Te /L, themaximum instability occurs approximately atxS /L5 1

2(4k13)/(2n11). For the lowest moden50, there is no integerk such thatxS /L<1. The instability condition cannot btherefore satisfied for the lowest mode of a tube open athot end.

C. Both ends open

For a tube open at both ends, with the cold endx50, thenth eigenfrequency is approximately given by~54!.The same relations~67! and~68! of the previous case applyThe condition for instability is again given by~58!. WhenudTw /dxu@Te /L the maximum instability occurs forxS /L5 1

4@(4k11)/n#(cC /ce), approximately. For the fundamental moden51 we thus havek50 andxS5 1

4(cC /ce)L.

IV. COMPARISON WITH EXACT THEORY

We compare here the results of the multiple-time-scacalculation and the short stack approximation with thosetained from the exact equation~11! for some typical param-eter values.

The geometry that we simulate is that of the experimeof Atchley7,8 in which the tube length was 99.87 cm, iradius 3.82 cm, and the stack lengthLS53.5 cm. The com-bined cross-sectional area of the stack plates was 3.12,i.e., 27% of the entire cross section. We have calculatedexact eigenvaluev both accounting for, and neglecting, thblockage effect of the plates. As the differences betweentwo cases are very small, in the results shown belowneglect blockage for simplicity.

Of course, our purpose here is not to compare Rotheory or our approximations with the data, a task thatalready been carried out, e.g., by Atchley himself aothers,7–9,19,10but to validate the asymptotic results againthe exact ones.

The gas used in the experiment was helium and accingly we takeg55/3, s50.71, cp55.2 kJ/kg K. Over thetemperature range of interest here, from 300 K to 700 K,fit thermal conductivity data20 by a linear function of tem-perature ask50.15113.22831024(T2300), with k in W/m K and T in K, which provides a better fit than a powelaw. We have included this effect in our calculation as tvalue of k determines the boundary layer thickness, atherefore the heat transfer parameters. The measuredperature was approximately constant and equal to its cand hot values to the left and right of the stack respectivand linear in the stack and therefore we takeT5TC for 0<x<xS2 1

2LS , T5TH for xS1 12LS<x<L, and

T5TC1x2~xS2 1

2LS!

LS~TH2TC!, ~69!

for xS2 12LS,x,xS1 1

2LS . HerexS denotes the position othe midpoint of the stack.

To solve Eq.~11! numerically we multiply byS anddiscretize by centered differences on an equispaced g

3315 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

he

e

e

t

s-

ts

e

ee

’ss

dt

d-

e

edm-ldy,

id.

This procedure requires the values of (12 f V)S at the half-integer nodes which are calculated as simple arithmeticerages. Typically 2000 cells were used along the tube lenwith approximately 100 in the stack region. The eigenvaluwere searched by the inverse iteration method.21 The samemethod was used to calculate the eigenvalues and eigenftions of the first-order approximation~21!. The integrationsnecessary to calculateV in Eqs.~33! and~34! were effectedby the trapezoidal rule. In order to avoid introducing an aditional parameter—the radius of the tube—we have taf V and f K to vanish outside the stack region.

As the operation of a thermoacoustic prime movercritically dependent on the imaginary part of the eigenfquency, we focus on this quantity. In Figs. 1 and 2 we shIm v as a function of the stack positionxS /L normalized bythe tube length. These figures are for an undisturbed presof 307 kPa, with a temperature difference of 368 K alongstack. The choice of this particular case is suggested byexperimental conditions of Atchleyet al.9,22 The frequencyRev/2p of the fundamental mode changes from about 5Hz to 590 Hz as the stack is moved from the hot to tmiddle of the tube. Within about 10%, these values are wpredicted by the formula~54! with n51, z51. Since thethermal penetration depth depends on frequency, the rdK / l is also a function of the stack’s position. When thstack is in the middle of the tube this ratio has the value 0in Fig. 1 and 0.17 in Fig. 2. As the stack is moved towardhot end, the frequency decreases and the penetration dcorrespondingly increases to 0.37 for Fig. 1 and 0.19 for F2.

In Figs. 1 and 2 the solid lines mark the results from tsolution of the exact Rott equation~11!, the dotted lines arethe two-time-scales approximation~32!, the dashed lines arethe short-stack approximation~57!, and the dash-dot lines arthe results obtained from Eq.~80! in Swift6 according to~41!. The zero is marked by the horizontal line; a negatvalue of Imv implies instability, and a positive one stabilityAgreement between the two-times approximation~32! andthe exact solution is excellent even at the relatively lavalue ofdK / l of Fig. 1. The difference is hardly noticeablethe smaller value of Fig. 2, and practically disappearseven larger gaps. The short-stack approximation~57! is notas accurate, but the error is less than about 15% even inworst case. Swift’s result is good for the wider gap caseFig. 2 but, as Fig. 1 shows, it rapidly deteriorates as thewidth decreases. It should be noted that, in plotting Swiresult, the values of pressure and velocity appearingSwift’s formula have been obtained from the exact solutof Rott’s equation.

Another important quantity is the critical temperatugradient, which is shown in Figs. 3 to 5 as a function ofdK/ lfor stack positions atxS /L50.594, 0.729, 0.864, respectively. Here the geometry and conditions are as in the preous figures except for the hot temperatureTH which is evalu-ated as follows. In the case of the exact Rott solutionTH isadjusted so as to achieve marginal stability conditions~i.e.,Im v50! with a fixed stack lengthLS , after which the criti-cal temperature gradient is calculated as (TH2TC)/LS . ~Re-call that the stack temperature is prescribed to be linea

3315S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 8: Linear thermoacoustic instability in the time domain

thm

fo

ngvethheek

sin

eheble

the

b-

sultons

forthevel-p-ut

avexi-redher

he-iremes.anaskwis-

ac-

ns-

m

o

edy

m

tavetio

ox

mal

tacke

tion

oxi-

Redistrib

this work.! These results are shown by the solid lines infigures. The short-stack approximation to the critical teperature gradient has been calculated from~51! which, withdTw /dx5(TH2TC)/LS , can be regarded as an equationTH . The dotted line shows this result withP1 approximatedby ~56!, while the dashed line is found from a correspondiapproximation,~66!, based on the cold-temperature wanumber. For the two positions closest to the hot end oftube~Figs. 4 and 5! most of the tube is occupied by gas at tcold temperature and use of the second form of the eigfunction gives a somewhat better result. When the stacpositioned near the midpoint of the tube~Fig. 3!, on the otherhand, both alternatives give a comparable error that canreduced by using the average of the two approximationP1 ~not shown!. We have repeated these calculations us

FIG. 3. Critical temperature gradient as a function of the ratio of the therpenetration lengthdK to the gap widthl for a stack positioned atxS /L50.594, i.e., near the midpoint of the tube. The solid line is the exact Rresult, the dotted line the short-stack approximation~51! with P1 approxi-mated by the hot-temperature wave function~56!, and the dashed line by thcold-temperature wave function~66!. The dash-dot line is the result founby setting Swift’s~Ref. 6! Eq. ~80! to 0. The large discrepancy is principalldue to the use of the approximation~10! for the f V,K’s in ~48!.

FIG. 4. Critical temperature gradient as a function of the ratio of the therpenetration lengthdK to the gap widthl for a stack positioned atxS /L50.729. The solid line is the exact Rott result, the dotted line the short-sapproximation~51! with P1 approximated by the hot-temperature wafunction ~56!, and the dashed line by the cold-temperature wave func~66!. The dash-dot line is the result found by setting Swift’s~Ref. 6! Eq.~80! to 0. The large discrepancy is principally due to the use of the apprmation ~10! for the f V,K’s in ~48!.

3316 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

e-

r

e

n-is

betog

the exact two-time-scale expression~51! with P1 found nu-merically from ~23!. When the stack is positioned near thtube’s midpoint the error is about 1%. When however tstack is moved toward the hot end, this result is comparato that obtained by the short-stack approximation withcold-temperature wave number.

In Figs. 3 to 5 the dash-dot lines show the result otained by setting Swift’s expression forW2 to 0. The errorincurred in the approximation off V,K by ~10! is seen to leadto a substantial discrepancy with the exact results. Our rewould show a comparable error if the same approximatifor f V,K were used.

V. CONCLUSIONS

By exploiting the difference between the time scalethe development of the thermoacoustic instability andperiod of standing waves in a resonant tube, we have deoped a time-domain description of the instability and aproximate formulae for its growth rate. We have carried oa multiple-time-scales expansion to first order and we hgiven simplified formulae exploiting the short-stack appromation. In addition to closed tubes, we have also considetubes open at either or both ends. Results for modes higthan the fundamental one have also been presented.

The exact perturbation results are very close to the toretical prediction of the full Rott equation. In practice, theevaluation requires the solution of the eigenvalue probl~21! that is self-adjoint and has therefore real eigenvaluAccordingly, numerically the problem is more tractable thRott’s equation. Even this rather modest computational tcan be avoided—at the price of an error of a fepercent—by adopting the short-stack approximation dcussed in Sec. III.

Another significant result of the present study is ancurate~and numerically manageable! approximation for thecritical temperature gradient including viscosity, heat tra

al

tt

al

ck

n

i-

FIG. 5. Critical temperature gradient as a function of the ratio of the therpenetration lengthdK to the gap widthl for a stack positioned atxS /L50.864. The solid line is the exact Rott result, the dotted line the short-sapproximation~51! with P1 approximated by the hot-temperature wavfunction ~56!, and the dashed line by the cold-temperature wave func~66!. The dash-dot line is the result found by setting Swift’s~Ref. 6! Eq.~80! to 0. The large discrepancy is principally due to the use of the apprmation ~10! for the f V,K’s in ~48!.

3316S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50

Page 9: Linear thermoacoustic instability in the time domain

hoha

bumncato

aitha

nul.

nd

th

d

ve

st.

arodel

-

,

ic

es

ry,

ar

Redistrib

fer, and cross-sectional area changes. As Figs. 3 to 5 sthe present expression is significantly more accurate tothers available in the literature.

As a final point, the approximate result~40! for the rateof energy loss of a thermoacoustic device below the instaity threshold enables one to readily calculate the minimpower required to maintain a certain temperature differeacross the stack, i.e., to operate the device as a refriger

Although we have only considered a particular typethermoacoustic devices, in which the heated region contthe so-called stack, the same method can be applied to oconfigurations. The present results imply that good accurmay be expected as a result of this approach.

ACKNOWLEDGMENTS

The authors would like to express their gratitude to oof the referees useful suggestions. They are also gratefthe Office of Naval Research for the support of this work

1N. Rott, ‘‘Damped and thermally driven acoustic oscillations in wide anarrow tubes,’’ Z. Angew. Math. Phys.20, 230–243~1969!.

2N. Rott, ‘‘Thermally driven acoustic oscillations, Part IV: Tubes wivariable cross section,’’ Z. Angew. Math. Phys.27, 197–224~1976!.

3N. Rott, ‘‘Thermoacoustics,’’ Adv. Appl. Mech.20, 135–175~1980!.4N. Rott, ‘‘Thermally driven acoustic oscillations, Part VI: Excitation anpower,’’ Z. Angew. Math. Phys.34, 609–626~1983!.

5J. Wheatley, ‘‘Intrinsically irreversible or natural heat engines,’’ inFron-tiers in Physical Acoustics, edited by D. Sette~North-Holland, New York,1986!, pp. 35–475.

6G. W. Swift, ‘‘Thermoacoustic engines,’’ J. Acoust. Soc. Am.84, 1145–1180 ~1988!.

7A. A. Atchley, ‘‘Standing wave analysis of a thermoacoustic prime mo

3317 J. Acoust. Soc. Am., Vol. 103, No. 6, June 1998

ution subject to ASA license or copyright; see http://acousticalsociety.org/c

w,n

il-

eor.fnser

cy

eto

r

below onset of self-oscillation,’’ J. Acoust. Soc. Am.92, 2907–2914~1992!.

8A. A. Atchley, ‘‘Analysis of the initial build-up of oscillations in a ther-moacoustic prime mover,’’ J. Acoust. Soc. Am.95, 1661–1664~1994!.

9A. A. Atchley, H. E. Bass, T. J. Hofler, and H.-T. Lin, ‘‘Study of athermoacoustic prime mover below onset of self-oscillation,’’ J. AcouSoc. Am.91, 734–743~1992!.

10M. Watanabe, A. Prosperetti, and H. Yuan, ‘‘A simplified model for lineand nonlinear processes in thermoacoustic prime movers. Part I. mand linear theory,’’ J. Acoust. Soc. Am.102, 3484–3496~1997!.

11J. Kevorkian and J. D. Cole,Perturbation Methods in Applied Mathematics ~Springer-Verlag, New York, 1996!, 2nd ed., Sec. 3.2.

12J. A. Murdock,Perturbations~Wiley, New York, 1991!, pp. 227–233.13E. J. Hinch,Perturbation Methods~Cambridge U.P., Cambridge, 1991!,

pp. 116–126.14P. M. Morse and H. Feshbach,Methods of Theoretical Physics~McGraw-

Hill, New York, 1953!.15A. W. Naylor and G. R. Sell,Linear Operator Theory in Engineering and

Science~Springer-Verlag, New York, 1982!, 2nd ed.16L. Landau and E. M. Lifshitz,Fluid Mechanics ~Pergamon, Oxford,

1959!.17A. D. Pierce,Acoustics~American Institute of Physics, Woodbury, NY

1989!, 2nd ed.18Lord Rayleigh, The Theory of Sound~Macmillan, London, 1896, Re-

printed by Dover, 1945!.19A. A. Atchley and F. M. Kuo, ‘‘Stability curves for a thermoacoust

prime mover,’’ J. Acoust. Soc. Am.95, 1401–1404~1994!.20N. B. Vargaftik, Handbook of Physical Properties of Liquids and Gas

~Wiley, New York, 1975!.21W. H. Press, W. T. Vetterling, S. A. Teukolsky, and B. P. Flanne

Numerical Recipes in FORTRAN~Cambridge U.P., Cambridge, 1992!,2nd ed.

22A. A. Atchley, H. E. Bass, and T. J. Hofler, ‘‘Development of nonlinewaves in a thermoacoustic prime mover,’’ inFrontiers in NonlinearAcoustics, edited by M. F. Hamilton and D. T. Blackstock~Elsevier, NewYork, 1990!, pp. 603–608.

3317S. Karpov and A. Prosperetti: Linear thermoacoustic instability

ontent/terms. Download to IP: 132.174.255.116 On: Sat, 20 Dec 2014 15:49:50