13
Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow A. Piccolo Department of Civil Engineering, University of Messina, Contrada di Dio, 98166 S. Agata (Messina), Italy article info Article history: Received 10 December 2010 Received in revised form 16 June 2011 Available online 6 July 2011 Keywords: Thermoacoustics Heat exchanger Heat Sound Finite differences abstract A simplified computational method for studying the heat transfer characteristics of parallel plate ther- moacoustic heat exchangers is presented. The model integrates the thermoacoustic equations of the stan- dard linear theory into an energy balance-based numerical calculus scheme. Details of the time-averaged temperature and heat flux density distributions within a representative domain of the heat exchangers and adjoining stack are given. The effect of operation conditions and geometrical parameters on the heat exchanger performance is investigated and main conclusions relevant for HX design are drawn as far as fin length, fin spacing, blockage ratio, gas and secondary fluid-side heat transfer coefficients are con- cerned. Most relevant is that the fin length and spacing affect in conjunction the heat exchanger behavior and have to be simultaneously optimized to minimize thermal losses localized at the HX-stack junctions. Model predictions fit experimental data found in literature within 36% and 49% respectively at moderate and high acoustic Reynolds numbers. Ó 2011 Elsevier Ltd. All rights reserved. 1. Introduction The enhanced use of renewable energies has nowadays become a major concern for the world energy policy to meet the target of sustainable development. Thermoacoustic (TA) technology can play a significant role in this context because of its remarkable advantages over conventional energetic technologies. TA engines are sound-heat energy conversion devices which operate with no moving components, use no-polluting working gases, can be powered by low-grade energetic inputs (waste heat, solar energy, etc.) and which join construction simplicity/reliability to low fabrication costs. The physical key mechanism for these favorable technical fea- tures is the ‘‘thermoacoustic effect’’ where a sound wave naturally constrains fluid particles oscillating near solid surfaces to undergo properly phased pressure changes and heat transfers through a thermodynamic cycle [1,2]. The sound pressure–velocity phase an- gle (imposed by an acoustic network) distinguishes between trav- eling-wave and standing-wave engines. In the former the thermal interaction of the oscillating gas with the solid boundaries of a regenerator (a high heat capacity porous medium of small hydrau- lic radius) results in Stirling-type cycles with engine conversion efficiencies above 40% of the Carnot efficiency [3]. In the latter the fluid particles perform Brayton-type cycles in a stack (of larger pores) but ‘‘intrinsic’’ thermal irreversibilities limit the engine per- formances typically below 20% of Carnot’s [3]. For the thermoacoustic effect to be fully exploitable in power- production/refrigeration applications the stack/regenerator must work in conjunction with a couple of heat exchangers (HXs) which transfer heat between its edges and external heat sources and sinks. These components have recently been the object of many re- search studies since a significant improvement of the overall en- gine’s performance is expected to derive from their optimization. The finned-tube [4], and shell and tube [5] HX configurations are commonly used but reliable and univocal design criteria are still lacking. The main goal in designing efficient HXs is the achieve- ment of high transfer rates in conjunction to low acoustic dissipa- tion. Addressing this task for oscillatory fluid flows entails a series of new challenges compared to traditional compact HXs working in steady and unidirectional flows: (a) Longitudinal (along the x direction parallel to the particle oscillation) and transverse (along the y direction normal to the fin/plate surfaces) dimensions are bounded respectively by the particle displacement amplitude x 1 and by the ther- mal penetration depth d j ð¼ ffiffiffiffiffiffiffiffiffiffiffiffiffi 2j=x p Þ, the distance through which heat can diffuse in an acoustic cycle (j and x being respectively the gas thermal diffusivity and the angular fre- quency of the sound wave). This imposes severe restrictions 0017-9310/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijheatmasstransfer.2011.06.027 Tel.: +39 090 3977311; fax: +39 090 3977480. E-mail address: [email protected] International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 Contents lists available at ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Embed Size (px)

Citation preview

Page 1: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer

journal homepage: www.elsevier .com/locate / i jhmt

Numerical computation for parallel plate thermoacoustic heat exchangersin standing wave oscillatory flow

A. Piccolo ⇑Department of Civil Engineering, University of Messina, Contrada di Dio, 98166 S. Agata (Messina), Italy

a r t i c l e i n f o a b s t r a c t

Article history:Received 10 December 2010Received in revised form 16 June 2011Available online 6 July 2011

Keywords:ThermoacousticsHeat exchangerHeatSoundFinite differences

0017-9310/$ - see front matter � 2011 Elsevier Ltd. Adoi:10.1016/j.ijheatmasstransfer.2011.06.027

⇑ Tel.: +39 090 3977311; fax: +39 090 3977480.E-mail address: [email protected]

A simplified computational method for studying the heat transfer characteristics of parallel plate ther-moacoustic heat exchangers is presented. The model integrates the thermoacoustic equations of the stan-dard linear theory into an energy balance-based numerical calculus scheme. Details of the time-averagedtemperature and heat flux density distributions within a representative domain of the heat exchangersand adjoining stack are given. The effect of operation conditions and geometrical parameters on the heatexchanger performance is investigated and main conclusions relevant for HX design are drawn as far asfin length, fin spacing, blockage ratio, gas and secondary fluid-side heat transfer coefficients are con-cerned. Most relevant is that the fin length and spacing affect in conjunction the heat exchanger behaviorand have to be simultaneously optimized to minimize thermal losses localized at the HX-stack junctions.Model predictions fit experimental data found in literature within 36% and 49% respectively at moderateand high acoustic Reynolds numbers.

� 2011 Elsevier Ltd. All rights reserved.

1. Introduction

The enhanced use of renewable energies has nowadays becomea major concern for the world energy policy to meet the target ofsustainable development. Thermoacoustic (TA) technology canplay a significant role in this context because of its remarkableadvantages over conventional energetic technologies. TA enginesare sound-heat energy conversion devices which operate with nomoving components, use no-polluting working gases, can bepowered by low-grade energetic inputs (waste heat, solar energy,etc.) and which join construction simplicity/reliability to lowfabrication costs.

The physical key mechanism for these favorable technical fea-tures is the ‘‘thermoacoustic effect’’ where a sound wave naturallyconstrains fluid particles oscillating near solid surfaces to undergoproperly phased pressure changes and heat transfers through athermodynamic cycle [1,2]. The sound pressure–velocity phase an-gle (imposed by an acoustic network) distinguishes between trav-eling-wave and standing-wave engines. In the former the thermalinteraction of the oscillating gas with the solid boundaries of aregenerator (a high heat capacity porous medium of small hydrau-lic radius) results in Stirling-type cycles with engine conversion

ll rights reserved.

efficiencies above 40% of the Carnot efficiency [3]. In the latterthe fluid particles perform Brayton-type cycles in a stack (of largerpores) but ‘‘intrinsic’’ thermal irreversibilities limit the engine per-formances typically below 20% of Carnot’s [3].

For the thermoacoustic effect to be fully exploitable in power-production/refrigeration applications the stack/regenerator mustwork in conjunction with a couple of heat exchangers (HXs) whichtransfer heat between its edges and external heat sources andsinks. These components have recently been the object of many re-search studies since a significant improvement of the overall en-gine’s performance is expected to derive from their optimization.The finned-tube [4], and shell and tube [5] HX configurations arecommonly used but reliable and univocal design criteria are stilllacking. The main goal in designing efficient HXs is the achieve-ment of high transfer rates in conjunction to low acoustic dissipa-tion. Addressing this task for oscillatory fluid flows entails a seriesof new challenges compared to traditional compact HXs working insteady and unidirectional flows:

(a) Longitudinal (along the x direction parallel to the particleoscillation) and transverse (along the y direction normal tothe fin/plate surfaces) dimensions are bounded respectivelyby the particle displacement amplitude x1 and by the ther-mal penetration depth djð¼

ffiffiffiffiffiffiffiffiffiffiffiffiffi2j=x

pÞ, the distance through

which heat can diffuse in an acoustic cycle (j and x beingrespectively the gas thermal diffusivity and the angular fre-quency of the sound wave). This imposes severe restrictions

Page 2: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Nomenclature

a speed of sound (m s�1)A surface area (m2)cP isobaric specific heat (J kg�1 K�1)cF specific heat of the secondary fluid (J kg�1 K�1)DR drive ratio (=PA/P0)_e energy flux density (W m�2)f spatially averaged h functionh thermoviscous function, specific enthalpy (J m�3), heat

transfer coefficient (W m�2 K�1)j imaginary unitjC Colburn-j factork wave number (m�1), thermoviscous function (m)K thermal conductivity (W m�1 K�1)l half plate/fin thickness (m)L heat exchanger or stack length (m)_m secondary fluid mass flux (kg s�1)

Mol gas molecular weight (kg kmol�1)Nu Nusselt numberp acoustic pressure (Pa)P pressure (Pa)Pr Prandtl numberPT internal perimeter of a tube section (m)_q time-averaged heat flux density (W m�2)_Q heat load (W)_Q� heat flux at the fin surface (W)R gas universal constant (J K�1 kmol�1)R0 R Mol�1 (kJ K�1 kg�1)t time (s)T temperature (K)u specific internal energy (J m�3)U fin-to-reservoir unitary global conductance

(W m�2 K�1)v acoustic particle velocity (m s�1)x axial coordinate/direction (m)xs stack center coordinate (m)y transverse coordinate/direction (m)y0 half plate/fin spacing (m)

Greek symbolsa fit constant of experimental values of KS, KC and KH

(W m�1 K�1)

b thermal expansion coefficient (K�1)c ratio of isobaric to isochoric specific heatd penetration depth, mg shear viscosity coefficient (N m�2 s�1)h TF-T0 (K)DTHXS stack-HX T difference (K)DTHXg gas-HX T difference (K)j gas thermal diffusivity (m2 s�1)k wavelength of the sound wave (m)l hFPT= _mcF

m kinematic viscosity (m2 s�1)n generic acoustic variableP thermoviscous functionq density of the gas (kg m�3)U viscosity stress tensor (N m�2 s�2)v fit constant of experimental values of KS, KC and KH (K�1)x angular frequency (rad s�1)X blockage ratio

Subscripts0 mean, time averaged1 first order acoustic variableA acoustic amplitude at a pressure antinodeC referred to the cold HXF secondary fluidH referred to the hot HXin evaluated at the inlet of the HXP isobaricres resonatorref reference values stacksolid solidx longitudinal, x-componenty transversal, y-componentz along the z directionj thermalm viscous

Superscriptsin evaluated at the inlet of the HX⁄ evaluated at the fin surface

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4519

on the heat transfer area and on the flow resistanceminimization.

(b) The junctions between the stack/regenerator and the HXsare interested by complex temperature and flow patternsdue to cross-sectional abrupt changes and entrance effects[6,7]. These phenomena may strongly affect the actual heattransfer and power dissipation rates.

(c) Unambiguous and reliable heat transfer correlations laws forthe oscillatory flow regime in the sub-kHz domain are notyet available. Conventional steady flow correlations arecommonly applied for the estimation of the heat transfercoefficients previous ‘‘ad-hoc’’ modifications [8,9].

(d) Large temperature differences are imposed by the hot andcold HXs to the ends of the stack/regenerator (typically longonly several centimeters). These high temperature gradientsimmediately results in performance degradation (especiallyin TA refrigerators) due to the associated increased thermalirreversibilities. Temperature differences between the HXsand the stack/regenerator should be minimized preservinghowever an efficient inter-element heat exchange.

These issues cannot be faced by the standard linear theory [1,2]owing its 1D character and being based on the so called ‘‘meanfield approximation’’ [10]. Thus 2D or 3D numerical and analyticalmodels able to account for the transverse heat transfer phenomenataking place in the HXs are currently developed [11–15]. In this pa-per the results of a 2D numerical model for the study of the ther-mal performance of parallel plate thermoacoustic HXs coupled toa stack in a simplified configuration are presented. The model de-scribes resonator acoustics through an idealized 1D sound wavewhile performs detailed 2D computations of the time-averagedtemperature and energy flux distributions within a representativedomain of the stack and HXs. These last are generated by an energybalance-based numerical calculus scheme which implements theenergy transport variables of the classical linear thermoacoustictheory. The model constitutes an extended version of the one pre-sented in [8,16] (simulating a thermally isolated stack) by inclu-sion of hydrodynamic heat transport along the transverse ydirection, temperature dependent thermophysical gas/solidparameters, viscous dissipation and adjacent HXs interacting with‘‘hot’’ and ‘‘cold’’ thermal reservoirs. The model is validated by

Page 3: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

4520 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

comparing numerical predictions against experimental data foundin literature. The effect of operating conditions and geometricalparameters on the heat transfer characteristics are investigated.Information on the optimal settings which maximize the HX per-formance is derived.

2. Formulation

The thermoacoustic model system considered in this work is astack of parallel plates, of length LS, sandwiched between ‘‘hot’’and ‘‘cold’’ parallel-fin HXs respectively of length LH and LC. Thisassembly is located at a distance xS from the center of a half-wave-length (k/2) gas filled resonator sustaining a standing acousticwave as shown in Fig. 1. Stack plates and HX fins are modeled withthe same spacing (2y0) and thickness (2l) so the three structuresresult characterized by the same blockage ratio X = As/Ares = 1/(1 + l/y0), where Ares is the cross sectional area of the resonatorand As is the cross sectional area of the stack/HXs open to gas flow.This choice, while allowing for a simple geometry, should lead to areduction of the detriment effects discussed in Section 1 at point b,being the flow and energy path almost not interrupted at the junc-tions. The implicit assumption is that an efficient HX is being mod-eled. The HX fins are set close to the ends of the stack plates but arenot in touch with them in order to reduce adverse heat conductionleaks from hot to cold HX. To include in the calculation the effect ofthe heat transfer coefficient from the secondary fluid-side, the HXsare modeled to thermally interact with two thermal reservoirs, the‘‘hot’’ reservoir at temperature TH and the ‘‘cold’’ reservoir at tem-perature TC. Since we are considering a thermoacoustic deviceworking in the refrigeration mode the former acts as a heat sinkwhile the latter acts as a heat source.

In the rest of this section the basic energy balance equations onwhich the model relies upon are defined together with the expres-sions of the energy transport variables entering in them. Prelimi-narily, the main assumptions involved in the model are discussed.

2.1. Basic approximations and simplifying conditions

The formulation relies on the following assumptions:

� The problem is two-dimensional;� The acoustic approximation is valid (low Mach number regime)

and any acoustic variable n can expresses in complex notationby the conventional first-order expansion nðx; y; tÞ ¼ n0ðx; yÞþRefn1ðx; yÞejxtg where t is the time, j the imaginary unit, Re{ }signifies the real part and where the mean value n0 is real butthe amplitude n1 is complex to account for both magnitudeand phase of the oscillation.� The working fluid is a Newtonian gas with viscosity obeying the

ideal gas state equation P0/q0 = R0T0 where P0 is the static pres-sure, q0 the static density, T0 the mean (time-averaged) temper-

Fig. 1. Schematic illustration of the modeled thermoacoustic refrigerator.

ature and R0 = R/Mol (R being the universal constant of gases andMol the molecular weight of the gas under study). Temperaturedependence of the gas thermophysical properties b (thermalexpansion coefficient), g (shear viscosity coefficient) and K(thermal conductivity coefficient) are accounted for by therelations

b ¼ 1T0; g ¼ gref

T0

Tref

� �0:7

; K ¼ KrefT0

Tref

� �0:7

ð1Þ

gref and Kref being the values of g and K at the reference temper-ature Tref .� The stack is significantly shorter than the acoustic wavelength

and is not intrusive. This allows the sound field to be retainedas spatially uniform over the HXs/stack assembly and to beapproximated in the neighbor (but also inside the stack as faras the pressure is concerned) by a 1D loseless standing wave:

p1 ¼ PA sinðkxsÞ ¼ p0; vx1 ¼ jPA

q0acosðkxsÞ ¼ jv0 ð2Þ

p1 being the amplitude of the dynamic pressure, PA the ampli-tude of the dynamic pressure at a pressure antinode, vx1 theamplitude of the longitudinal particle velocity, k the wave num-ber (k = 2p/k) and a the sound velocity.� The standard thermoacoustic equations are considered for the

amplitudes of the oscillating temperature T1, longitudinal veloc-ity vx1, transverse velocity gradient ovx1/oy transverse velocityvy1 and transverse velocity gradient ovy1/oy, inside a gas pore.If the specific heats of the plate/fin materials are notably greaterthan the isobaric specific heat of the gas cp (negligible temper-ature oscillations in the solid), the first three equations read as[1,2]:

T1 ¼1

q0cPð1� hjÞp1 �

1q0x2ð1� PrÞ

dp1

dx@T0

@x½ð1� hjÞ

� Prð1� hmÞ� ð3Þ

vx1 ¼j

xq0

dp1

dxð1� hmÞ;

@vx1

@y¼ 2

xq0d2m

dp1

dxkm ð4Þ

where

hj ¼cosh½ð1þ jÞy=dj�cosh½ð1þ jÞy0=dj�

; hm ¼cosh½ð1þ jÞy=dm�cosh½ð1þ jÞy0=dm�

and

kj ¼dj

1þ jsinh½ð1þ jÞy=dj�

cosh½ð1þ jÞy0=dj�; km ¼

dm

1þ jsinh½ð1þ jÞy=dm�

cosh½ð1þ jÞy0=dm�

and where Pr is the Prandtl number, dm ¼ffiffiffiffiffiffiffiffiffiffiffiffi2m=x

pthe viscous pen-

etration depth (m being the kinematic viscosity) and where y = 0 isfixed at the center of the fluid gap.

The analytical expression of vy1 can be derived from the linearizedequation of continuity

jxq1 þ@

@xðq0vx1Þ þ

@

@yðq0vy1Þ ¼ 0 ð5Þ

Neglecting the term (oq0/oy)vy1 in comparison to the term (oq0/ox)vx1, (vy1 being expected to be of order dm/k smaller than vx1)and taking into account the relations

1q0

@q0

@x¼ � 1

T0

@T0

@x; q1 ¼ �q0bT1 þ

ca2 p

(c being the ratio of isobaric to isochoric specific heat) Eq. (5) can beput into the form

Page 4: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Fig. 2. Profile of the transverse velocity amplitude vy1 with coordinate y.

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4521

@vy1

@y¼ jxbT1 � j

cxq0a2 p1 �

@vx1

@xþ b

@T0

@xvx1 ð6Þ

Eqs. (3) and (4) can now be substituted into Eq. (6) for T1 and vx1

while the axial velocity gradient can be deduced from the waveequation (Eq. (A19) of Ref. [1]). Making substitutions and rearrang-ing (by also using the thermodynamic relation cp = ba2/(c � 1)) weobtain

Table 1Parameters of selected simulations. In all runs test-gas = He, P0 = 101325 Pa, mean tedj = 5.352 � 10�4 m, TH = 300 K, TC = 297 K, LS = 0.07 m.

Run y0/dj l/dj xs/k PA/P0 (%)

1 3 0.93 0.11 4.932 1.5 0.47 0.11 4.933 1.5 0.47 0.11 0.49–5.434 1.5 0.47 0.11 0.49–5.435 1.5 0.47 0.11 0.49–5.436 1.5 0.47 0.11 3.957 1.5 0.47 0.11 4.448 1.5 0.47 0.11 4.939 1.5 0.47 0.11 5.43

10 1.5 0.47 0.11 2.4711 1.5 0.47 0.11 3.4512 3 0.93 0.11 0.49–5.9213 1.5 0.47 0.11 0.49–5.9214 1.5 0.47 0.11 0.49–7.9015 1.5 0.47 0.11 2.9616 1.5 0.47 0.11 4.9317 1.5 0.47 0.11 6.9118 0.5–3 0.16–0.93 0.11 4.9319 0.5–3 0.16–0.93 0.11 4.9320 0.5–3 0.16–0.93 0.11 4.9321 0.5–3 0.16–0.93 0.11 4.9322 0.5–3 0.16–0.93 0.11 4.9323 0.5–3 0.16–0.93 0.11 4.9324 2 0.8 0.11 1.6325 2 0.8 0.11 6.9126 0.5–3 0.75 0.11 4.9327 0.5–3 0.75 0.11 4.9328 0.5–3 0.75 0.11 4.9329 1.5 0.47 0.09 7.9–8.930 1.5 0.47 0.09–0.08 8.931 1.5 0.47 0.07 9.932 1.5 0.47 0.11 0.99–7.8933 1.5 0.47 0.1–0.09 7.8934 2.0 0.8 0.11 0.69–8.035 2.0 0.8 0.11–0.1 8.036 2.0 0.8 0.09–0.07 8.037 2.0 0.8 0.07 8.4–10.8

@vy1

@y¼ j

xq0a2

½1þðc�1Þfj�ð1�hmÞ� ½1þðc�1Þhj�ð1� fmÞð1� fmÞ

� �p1

þ jb

q0xð1�PrÞfmð1�hjÞ� fjð1�hmÞþ ðhj�hmÞ

ð1� fmÞ

� �@T0

@xdp1

dxð7Þ

whose integration between 0 and y provides

vy1 ¼ jx

q0a2

½1þðc�1Þfj�ðy�kmÞ� ½yþðc�1Þkj�ð1� fmÞð1� fmÞ

� �p1

þ jb

q0xð1�PrÞfmðy�kjÞ� fjðy�kmÞþðkj�kmÞ

ð1� fmÞ

� �@T0

@xdp1

dxð8Þ

where

fj ¼tanh½ð1þ jÞy0=dj�½ð1þ jÞy0=dj�

; f m ¼tanh½ð1þ jÞy0=dm�½ð1þ jÞy0=dm�

In Fig. 2 the profile of the real and imaginary parts of Eq. (8) to-gether with its module is reported as a function of the transversecoordinate y for parameters corresponding to run 1 of Table 1.The transverse velocity vanishes both at mid channel y = 0 (for sym-metry) and at the wall surface y = y0.� The longitudinal pressure gradient dp1/dx along the stack/HXs

assembly is calculated by imposing continuity of volume flowrate at the entrance of the HXs

mperatu

dp1

dx¼ q0xð1� fmÞX

v0 ð9Þ

We remark as the validity of the above derived equations could beinvalidated at the pore ends of the stack or at the HX fins where the

re = 300 K, resonance frequency = 200 Hz, Uref = 100 W m2 K�1, k = 10.08 m,

LC/LS LH/LS UC/Uref UH/Uref

0.17 0.18 1 10.11 0.11 0.1 0.10.11 0.11 0.1 0.10.11 0.11 1 10.11 0.11 5 50.11 0.11 0.1–20 0.1–200.11 0.11 0.1–20 0.1–200.11 0.11 0.1–20 0.1–200.11 0.11 0.1–20 0.1–200.11 0.11 0.1–20 0.1–200.11 0.11 0.1–20 0.1–200.11 0.11 10 100.11 0.11 20 200.11 0.11 20 200.0143–0.343 0.18 30 300.0143–0.343 0.18 30 300.0143–0.343 0.18 30 300.0143 0.18 30 300.036 0.18 30 300.071 0.18 30 300.126 0.18 30 300.180 0.18 30 300.343 0.18 30 300.14 0.14 20 200.14 0.14 20 200.036 0.18 30 300.018 0.18 30 300.343 0.18 30 300.29 0.29 20 200.29 0.29 20 200.29 0.29 20 200.11 0.11 20 200.11 0.11 20 200.14 0.14 20 200.14 0.14 20 200.43 0.43 20 200.43 0.43 20 20

Page 5: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

4522 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

gas mean temperature is expected to become strongly y-dependentto allow net heat exchange with the solid surface. Eq. (3), instead,was derived as solution of the heat transport equation involving ay-independent T0. Analogously, invoking validity of Eqs. (4), (7)and (8) at the fin outer ends (facing the duct) corresponds to neglectany entrance effect which may be the source of minor losses. Re-sults of the present study have thus to be appreciated as a functionof these major limitations.

2.2. Energy balance equation and energy fluxes in the gas

Under the simplifying conditions outlined in the previous sub-section the gas flow dynamics remains completely describedthrough Eqs. 2, 4, and 9. Meanwhile, the temperature field is gov-erned by the general equation of energy conservation [17]

@

@t12qv2 þ qu

� �¼ �r � _e ð10Þ

where u is the specific internal energy and where _e, the energy fluxdensity, is defined as

_e ¼ qv12

v2 þ h� �

� v �U� KrT ð11Þ

h being the specific enthalpy and U the viscous stress tensor. Timeaveraging Eq. (10) over an acoustic cycle leads to

r � �_e ¼ 0 ð12Þ

where over-bar means time-averaged. Eq. (12) is applied in thiswork to each computational cell of the simulation domain to im-pose local energy balance in the gas. To accomplish this, a finite dif-ference technique is employed where the quantitative results ofstandard linear theory are considered for x and y the componentsof the time-averaged energy flux density. These last are derived inthe following subsections.

2.2.1. Energy flux density along the x directionThe x component of the time-averaged enthalpy flux density is

[17]

�_ex ¼x2p

Z 2p=x

0qvx

12

v2 þ h� �

dt �Z 2p=x

0ðvxrxx þ vysxyÞdt

��K

Z 2p=x

0

@T@x

dt�

ð13Þ

where the components of the viscous stress tensor are

rxx ¼23g 2

@vx

@x� @vy

@y

� �� �2

3g@vy

@y

sxy ¼ syx ¼ g@vx

@yþ @vy

@x

� �� g

@vx

@y

ryy ¼23g 2

@vy

@y� @vx

@x

� �� 4

3g@vy

@y

the x velocity gradients having being neglected because of order dm/k smaller than the y velocity gradients. Although the second integralon the right hand side of Eq. (13) is generally neglected becauseincluding terms of second order in velocity, it is retained in thiswork because we are trying to identify the impact of loss mecha-nisms on the performance of the two HXs. Making use of the com-plex notation and retaining only terms up to second order Eq. (13)can be written as:

�_ex ¼12q0cPRe T1 ~vx1f g þ 1

3gRe

@vy1

@y~vx1

� �� 1

2gRe

@~vx1

@yvy1

� �� K

@T0

@x

ð14Þ

where tilde indicates complex conjugation. Substituting now Eqs.(2) and (9) into Eqs. (3), (4), (7) and (8) and these last in Eq. (14),the following expression is found, at second order in the acousticoscillation amplitude, for �_ex:

�_ex ¼1

2XIm

P1~P2

ð1�~f mÞ

( )P0v0�

q0cP

2xX2ð1�PrÞj1� fmj2@T0

@xIm P1

~P2

n ov2

0

þ gx3Xq0a2 Re

~P2P5

ð1�~f mÞ

( )p0v0þ

gb

3X2ð1�PrÞj1� fmj2

�@T0

@xRef eP2P6gv2

0þgx

q0Xa2d2m

Im~kmP3

ð1�~f mÞ

( )p0v0

þ gb

d2mX

2ð1�PrÞj1� fmj2@T0

@xImf~kmP4gv2

0�K@T0

@xð15Þ

where

P1 ¼ ð1� hjÞ

P2 ¼ ð1� hmÞ

P3 ¼½1þ ðc� 1Þfj�ðy� kmÞ � ½yþ ðc� 1Þkj�ð1� fmÞ

ð1� fmÞ

P4 ¼fmðy� kjÞ � fjðy� kmÞ þ ðkj � kmÞ

ð1� fmÞ

P5 ¼½1þ ðc� 1Þfj�ð1� hmÞ � ½1þ ðc� 1Þhj�ð1� fmÞ

ð1� fmÞ

P6 ¼fmð1� hjÞ � fjð1� hmÞ þ ðhj � hmÞ

ð1� fmÞ

2.2.2. Energy flux density along the y directionThe time averaged hydrodynamic enthalpy flux along the trans-

verse direction is

�_ey ¼x2p

Z 2p=x

0qvyð

12

v2 þ hÞdt �Z 2p=x

0ðvxsxy þ vyryyÞdt

��K

Z 2p=x

0

@T@y

dt�

Proceeding analogously as done before for �_ex we obtain

�_ey ¼12q0cPRefT1 ~vy1g �

12gRe

@vx1

@y~vx1

� �� 2

3gRe

@vy1

@y~vy1

� �� K

@T0

@y

ð16Þ

which, by substitution of the explicit expression of the acousticvariables, becomes

�_ey¼x

2q0a2 ImfP1eP3gp2

0þb

2Xð1�PrÞ ImP1eP4

ð1�~f mÞ

( )@T0

@xp0v0

þ cp

2Xð1�PrÞa2 Re jðP1�PrP2Þ eP3

ð1� fmÞ

( )@T0

@xp0v0

þ bq0cp

2xX2ð1�PrÞ2j1� fmj2RefjðP1�PrP2Þ eP4g

@T0

@x

� �2

v20

� gX2d2

m j1� fmj2Im km eP2

n ov2

0�2gx2

3q20a4

Ref eP3P5gp20

� 2gbx3q0Xa2ð1�PrÞ

@T0

@xRe

eP4P5

ð1�~f mÞþeP3P6

ð1� fmÞ

( )p0v0

� 2gb2

3X2ð1�PrÞ2j1� fmj2@T0

@x

� �2

Ref eP4P6gv20�K

@T0

@yð17Þ

Page 6: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4523

Note as in the expressions derived for �_ex and �_ey the dependence onthe stack position inside the resonator is provided by p0 and v0. Inthe rest of this study these energy flux densities are identified toheat flux densities (�_ex _qx;

�_ey _qy) because in the short stackapproximation the work flux contribution to the total energy fluxis small [1] (the over-bars having been neglected to simplify thenotation).

2.3. Energy balance equation and energy fluxes in the solid

The energy equations in the gas are coupled to the ones in thesolid through heat exchange at the solid surfaces. The analogueof Eq. (12) for the solid portions is

r � _q ¼ 0 ð18Þ

where

_qx ¼ �KS@T0

@x; _qx ¼ �KH

@T0

@x; _qx ¼ �KC

@T0

@xð19Þ

and

_qy ¼ �KS@T0

@y; _qy ¼ �KH

@T0

@y; _qy ¼ �KC

@T0

@yð20Þ

KS being is the thermal conductivity of the plate material and KH andKC the thermal conductivities respectively of the ‘‘hot’’ and ‘‘cold’’HXs. The temperature dependence of these parameters is describedby empirical laws of the form Ksolid = Kref + a exp(±vT0), with Kref, aand v determined from the fit of experimental data found inliterature.

2.4. Energy balance in the heat exchangers

The cooling power _Q C produced thermoacoustically in thestack is supposed to be extracted through the cold HX from a coldthermal reservoir at temperature TC. Analogously, the sum of _QC

and of the heat produced by viscous dissipation over the surfacesof the stack/HXs assembly is rejected through the hot HX to a hotthermal reservoir at temperature TH. The modeled HXs are of theflat tube type (as those considered in [18]) with secondary fluidsflowing inside the tubes along the z direction (see Fig. 3). Thisconfiguration leads to a more simplified modeling compared toother geometries but introduces in the problem a z dependencenot included in our 2-D (x, y) treatment. To overcome this limita-tion we assume for the inner surface of the HXs tubing a constanttemperature wall boundary condition (compatible with the highthermal conductivities of the HX material) but substitute for thelocal heat flux density normal to the inner tubing surface _qyðzÞ aproper mean value obtained by averaging _qyðzÞ over the lengthLZ of the tube. We start from the well-known expression of thetemperature variation TF(z) of a fluid along a channel with con-stant wall temperature T0

hðzÞ ¼ hin expð�lzÞ

where hðzÞ ¼ TFðzÞ � T0; hin ¼ TinF � T0 and l ¼ hFPT= _mcF (Tin

F beingthe inlet fluid temperature, PT the internal cross section perimeter

Fig. 3. Schematics of the HXs and stack configuration.

of the tube, hF the convective coefficient fluid-wall, _m the mass fluxand cF the fluid specific heat). The total heat flux exchanged by thesecondary fluid is

_Q ¼Z Lz

0hF PThðzÞdz ¼ _mcF ½1� expð�lLZÞ�hin

So, the average heat flux density normal to the internal surface ofthe flat tube can be expressed in the form

_qy ¼_Q

PT LZ¼ UðTC � T0Þ ð21Þ

where

U ¼_mcF

PT LZ½1� expð�lLZÞ�

plays the role of a fin-to-reservoir unitary global conductance if Tin

is identified with TC or TH (in the hypothesis that the secondary fluidmake perfect thermal contact with the thermal reservoirs). Eq. (21)is used in this work to couple the heat flux thermoacoustically gen-erated in the stack to the heat carried out by the secondary fluid.

2.5. Numerical scheme

The periodicity of the stack and HXs structures along y allowscalculations to be performed in a single channel of the stack/HXsassembly and in the couple of plates/fins enclosing it. The compu-tational domain is further reduced by symmetry from half a gasduct to half a plate/fin as indicated in Fig. 4 by the light gray areatogether with the coordinate system used. The axis parallel to theplates is the x axis; x = 0 is chosen to be the beginning of the coldHX on the left. The y axis is perpendicular to the stack-plates; y = 0is chosen to be the midpoint between the two adjacent plates orfins. The calculation of the steady-state two-dimensional time-averaged temperature distribution is performed using a finitedifference methodology where temperature spatial gradients arediscretized using first order nodal temperature differences. To thisend, the computational domain is subdivided using a rectangulargrid. In the x direction the computation mesh size is typically0.005 Ls while in the y direction the computation mesh size is typ-ically 0.02 y0. The following boundary conditions are imposed:

� Symmetry on nodal lines at y = 0 and y = y0 + l

@T0

@y

� �y¼0¼ @T0

@y

� �y¼y0þl

¼ 0 ð22Þ

� continuity of temperature and transverse heat fluxes at the gas–solid interfaces (y = y0)

Fig. 4. Magnified view of the computational domain (gray area).

Page 7: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

4524 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

T0ðy ¼ y0Þ ¼ T0jsolidðy ¼ y0Þ; K@T0

@y

� �y¼y0

¼ Ksolid@T0

@y

� �y¼y0

ð23Þ

where subscript ‘‘solid’’ refers in general both to the stack plateand the HX fins.

� continuity of transverse heat fluxes at the inner surface of theHX tubing:

KCð@T0

@yÞtube ¼ UCðTC � T0Þ � KHð

@T0

@yÞtube ¼ UHðTH � T0Þ

ð24Þ

� Vanishing heat flux at the solid HX fin terminations(y0 6 y 6 y0 + l) facing the duct

@T0

@x

� �x¼0¼ @T0

@x

� �x¼ðLCþLSþLHÞ

¼ 0 ð25Þ

� Vanishing flux at the HX pore ends (0 6 y < y0) facing the duct(gas oscillations outside the HXs are adiabatic and thus thethermoacoustic effect vanishes)

_qxðx ¼ 0Þ ¼ _qxðx ¼ LC þ LS þ LHÞ ¼ 0 ð26Þ

Fig. 5. Centerline plate and fins temperature as a function of coordinate x. In theinsert the temperature discontinuities between the cold fin and the stack edge isreported as a function of drive ratio at selected U values.

As for the separation gap between the HX fin and the stack edge, itis set equal to dj and, as a further simplifying hypothesis, the sliceof gas enclosed in it is assumed at rest.

Substituting Eqs. (15) and (17) in Eqs. 12, 19, and 20 in Eq. (18)to impose local energy balance in each cell of the computationalgrid, a system of quadratic algebraic equations in the unknownvariable T0 is generated. The system is solved by a code developedby the author in FORTRAN-90 language which executes the recur-sive Newton–Raphson method [19]. At each iteration the system oflinear algebraic equations providing the temperature correctionsfor the next step is solved using a LU decomposition with partialpivoting and row interchanges matrix factorization routine. Thelatter is taken from the LAPACK library routines available onlineat [20], where details about accuracy, computation cost, etc. canbe found. Once the time-averaged temperature distribution isknown, it can be substituted in Eqs. (19) and (20) to determinethe energy flux distributions along the x and y directions both inthe gas and in the plate/fins. In addition, it can be used to calculateother variables of interest such as the energy dissipation, the en-tropy generation, etc.

Compared to other approaches currently used to simulatethermoacoustic devices the proposed model has the advantage ofa reduced computational cost since it directly deals with time-averaged quantities and does not require integration of the conser-vation laws of mass, momentum and energy over the large numberof acoustic cycles driving the system towards steady conditions. InSection 3.5 the code is validated by comparing numerical predic-tions against experimental data found in literature.

3. Results and discussion

The performance of the coupled stack-HXs system is investi-gated at varying both operating conditions (DR = PA/P0, xS, UH, UC)and stack/fin geometry (y0, l, LH, LC). The parameters of differentnumerical simulations (runs) are listed in Table 1. In all tests he-lium at atmospheric pressure and at a mean temperature of300 K is assumed as working fluid (m = 122 � 10�6 m2 s�1,K = 0.152 W m�1 K�1, a = 1008 m s�1, Pr = 0.68 at Tref = 300 K). It isconsidered enclosed in a half-wavelength resonator (acousticwavelength k = 5.04 m) having a resonance frequency of 200 Hz;The stack is 0.07 m long and is constituted by stainless steel(KS = 14.9 W m�1 K�1 at Tref = 300 K) while the HXs material is cop-

per (KC = KH = 401 W m�1 K�1 at Tref = 300 K). In all cases the tem-perature of the cold and hot reservoirs are fixed respectively toTC = 297 K and TH = 300 K. The dynamic Reynolds number (vx1dm/m) and Mach number (PA/q0a2) resulted always smaller respectivelythan 500 and 0.1, the values over which turbulence an non-linearbehavior are expected to affect fluid dynamics [1,21].

The cooling power, _Q C , is calculated by integrating _qy over theinner tubing surface of the cold HX

_QC ¼Z LC

0UC ½TC � T0ðxÞ�dx ð27Þ

with LZ arbitrarily set to 1 m. This quantity differs from the heat fluxentering/leaving the outer fin surface

_Q �C ¼Z LC

0K

@T0

@y

� �y0

dx ð28Þ

by the heat flux transferred through the vertical fin surface facingthe stack edge. Analogous quantities are calculated at the hot HXfin. At the operating conditions considered in this work _QC resultedsmaller than _Q �C typically by 3%, evidencing as the stack perfor-mance is degraded by the heat conduction losses at the stack-HXjunctions.

3.1. Effect of temperature difference

Fig. 5 illustrates the longitudinal temperature profile evaluatedin correspondence of the centerline (y = y0 + l) of the plate and ofboth fins for a drive ratio DR (=PA/P0) = 4.93% (run 2). The temper-ature distribution is linear in the stack (except at the stack edges)becoming flat in the fins, consistently with their high thermal con-ductivity (near isothermal fins). Temperature discontinuities DTHXS

are observed between the fins and the stack edges (with the stacktemperature higher than the cold fin temperature) in accordancewith experiments [22]. At low U values (U < 100) DTHXS increasesat increasing drive ratios as shown in the insert of the same figure(run 3). This behavior can be interpreted observing that in the pres-ent model the HXs are thermally interacting with hot and cold res-ervoirs. So, depending on the magnitude of the fin-to-reservoiroverall heat transfer coefficient U, the cold fin must cool down be-low TC and the hot fin must heat up above TH until the heat flux ex-tracted from the cold reservoir and the one delivered to the hotreservoir balance the heat flux thermoacoustically pumped along

Page 8: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Fig. 7. The cooling load as a function of the fin-to-reservoir unitary globalconductance at selected drive ratios. Continuous lines are guides for the eye.

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4525

the stack. At low U values this causes an increase of the mean lon-gitudinal temperature gradient across the HXs-stack assembly andthus of the temperature discontinuities observed at the HXs-stackjunctions. The same figure shows as the effect becomes lessmarked at increasing U values (run 4) till to invert its trend (DTHXS

decreases with DR) for U > 100 (run 5). In this latter case, in fact,the temperatures of the hot and cold HXs are ‘‘anchored’’ respec-tively to TH and TC and the enhanced hydrodynamic enthalpy fluxactivated by increasing DR are compatible with lower axial tem-perature gradients. This is clearly shown in Fig. 6 where the coolingload results to be a decreasing (linear) function of the ratio be-tween the temperature of the cold HX and the temperature ofthe hot HX (runs 6–9). Fig. 5 also reveals that at high U valuesDTHXS vanishes in correspondence of a given drive ratio and thenbecomes negative (stack temperature lower than fin temperature).This circumstance, also experimentally evidenced by Mozurkewich[9], could allow for minimization of the detriment heat leakagephenomenon from hot to cold HX discussed in the previous sub-section. The U magnitude, therefore, directly affects the tempera-ture difference across the stack and this in turn influences the axialhydrodynamic enthalpy flux, i.e. ultimately, the cooling load. HighU values are desirable to improve the performance of the device, asevidenced in Fig. 7 where the growth of the cooling load with U atselected DR is observable (runs 7, 9, 10, 11).

The existence of inter-element heat transfer under zero temper-ature difference is not surprising because the driving temperaturedifference for heat extraction/immission at the HXs is the one be-tween the HXs and the adjacent gas. High values of this transversetemperature gradient should be avoided, however, because theycould enhance the thermal irreversibility associated to the heattransfer process and, in addition, could determine a steep axialtemperature gradient in the gas which, as argued before, reducesthe useful cooling load. In Fig. 8 the mean gas-fin temperature dif-ference DTHXg = hTHXCi � hTgCi at the cold HX is plotted as a functionof DR at selected y0 values (runs 12, 13). Quantities hTHXCi and hTgCirepresent the spatially averaged temperatures respectively of thecold HX (evaluated at the gas-solid interface y = y0) and of thegas (evaluated at the centerline of the gas channel y = 0):

hTHXCi ¼1LC

Z LC

0T0ðx; y0Þdx; hTgCi ¼

1LC

Z LC

0T0ðx;0Þdx ð29Þ

The plot clearly shows the existence of two regimes: one at low DRwhere DTHXg exhibits a squared dependence on DR (DTHXg / DR2)

Fig. 6. The cooling load as a function of the ratio between the temperature of thecold HX and the temperature of the hot HX at selected drive ratios. Straight linesresult from a linear fit of the simulation points.

and one at high DR where DTHXg depends linearly on DR(DTHXg / DR). This result seems to conflict with the statements ofSwift [23] which predicts rather a linear behavior at low DR and asquared-one at high DR. The disagreement can be explained analyz-ing the Swift’s predictive formula based of the ’’thermal boundary-layer’’ theory [23]:

DTHXg ¼_Q Cdj

KAð30Þ

A (=LCLZ) being the heat transfer surface area. At low DR (for2x1 LC) it is expected _QC / DR2 so if 2x1 (/DR) is substituted forLC in A (as done in [23]) the linear behavior is found while if A isset always constant to LCLZ (as done in the present work) the qua-dratic behavior is obtained. At high DR (for 2x1� LC) if the func-tional relationship _QC / DR2 is retained (as done in [23]) and A isset equal to LCLZ the quadratic behavior is found but if the depen-dence of _QC on DR is not quadratic a deviation from the attendedbehavior should be expected. This is indeed the case of the presentstudy where simulations predict in the high drive ratio regime

Fig. 8. The mean temperature difference between the gas and cold HX fin as afunction of DR for two operating conditions. Dotted and continuous lines resultrespectively from a quadratic (low DR) and linear (high DR) fit of the simulationpoints.

Page 9: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

4526 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

(when 2x1� LC) a roughly linear dependence of _QC on DR (see nextsub-section).

An additional important information that can be drawn fromFig. 8 is that the choice y0/dj = 1.5 in conjunction withU = 2000 W m�2 K�1 entails lower transverse temperature gradi-ents and higher cooling loads (results not shown) compared tothe other. So for a given HX length, the benefit of reduced thermalirreversibilities could be met in conjunction with enhanced coolingloads for suitable HX configurations.

3.2. Effect of drive ratio

To investigate the effect of drive ratio on the cooling load _Q C

numerical simulations are performed at fixed fin lengths LC andLH. Results are shown in Fig. 9 where the mean heat loads per unitfin area _qC and _qH (and their difference in the insert) are plotted asa function of DR (run 15). The simulations reveal that at low driveratios (2x1 LC, LH) both _qC and _qH increase proportionally to DR2

as stated by the linear theory. At high drive ratios (2x1� LC, LH),however, the behavior of the heat loads differentiates: while _qH

continues to grow faster than linearly (even if with a reducedstrength), _qC exhibits a roughly linear dependence on DR. As al-ready pointed out by Swift [23], a candidate cause for the appear-ance of the linear behavior could be the inefficiency of the HXs intransferring the full thermal load at high pressure amplitudes. Thisshould happen when the peak to peak particle displacement ampli-tude exceeds the length of the heat exchangers. Although to theknowledge of the author there is no clear experimental evidencefor this transition effect in literature, some analogue findings arereported in the numerical studies of Worlikar and Knio [12] andof Marx and Blanc-Bennon [24]. The last authors, in particular, ob-served the same transition in behavior in the acoustically stimu-lated hydrodynamic energy flux flowing within the gas pores ofan isolated stack (without heat exchangers). It is interesting to noteas the effect (attributed by the authors to thermal wave anharmo-nicity) appeared at the drive ratio for which the peak-to-peak par-ticle displacement amplitude reached half the plate length. Thiscondition could be retained as equivalent to the analogue put for-ward here for the HXs when considering that the plate edges of anisolated stack act as real HX fins and that at the operation point ofthe transition the stack may be considered as constituted by twoHX fins (each of length LS/2) set in close contiguity.

As reported above, the analogous approaching of a linear behav-ior is not as evident in _qH as in _qC . This fact could be explained

Fig. 9. The heat loads per unit area on the cold and hot HX as a function of DR. In theinsert their difference is plotted. Dotted and continuous lines result respectivelyfrom a quadratic (low DR) and linear (high DR) fit of the simulation points.

observing that the hot HX have to sustain, in addition to the heatflux thermoacoustically pumped within the stack, the heat gener-ated by viscous dissipation over all the solid surfaces of the chan-nel (both stack and HXs). This contribute, which accounts for thelarger heat transfer area of the hot HX compared to the cold one,has a well-known quadratic dependence on DR and should corre-spond in our simplified model to the difference _qH � _qC . This lastis plotted in the insert of Fig. 9 where the quadratic dependenceon DR is clearly observable.

3.3. Effect of HX length and fin interspacing

The HX length along the axial direction is a critical parameter inHX design because it determines the available heat transfer areaand also accounts for the rate of viscous dissipation. To study theeffect of the cold HX length LC on the cooling load _Q C , simulationsare performed at fixed stack position and blockage ratio and atvarying DR, y0, LC and LH values. A typical result is shown inFig. 10 where _QC is plotted as a function of LC at selected DR values(runs 15–17). Simulations reveal that _QC increases quickly in therange 0 < LC < x1 becoming almost constant for LC > x1. At the con-sidered operating conditions and for LC/2x1 � 1 the cooling loadamounts at about 93% of the maximum cooling load evaluatedby calculating the total longitudinal enthalpy flux at the midpointof the stack (a choice justified by the fact that the stack plate islonger than the particle displacement for all the DR values consid-ered). As pointed out by Swift [1], this should be the optimal con-dition for efficient inter-element heat transfer. This conclusion isbased on the argument that a particle can transport heat over a dis-tance not exceeding 2x1 and that the total amount of particles ableto participate to the inter-element heat transfer corresponds to thegas portion comprised within a distance 2x1 from the stack edge.We remarkably note as for all the considered drive ratios halvingthe plate length starting from the above invoked optimal value(2x1 ? x1) produces a reduction of _Q C of less than 1% (from about93% to 92%). This result indicates, as already highlighted by Hofler[25], that the fin length could be chosen substantially shorter thanthe peak-to-peak particle displacement amplitude without com-promising the HX performance and with great benefits in termsof viscous losses reduction. An experimental evidence for this find-ing can be found in [23] where it was recognized that doubling thelength of the cold HX of a TA engine did not enhanced perfor-mance, even when the displacement amplitude much exceededthe length of the single HX.

Fig. 10. The cooling load as a function of the length of the cold HX at selected driveratios. Continuous lines are guides for the eye.

Page 10: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Fig. 12. The time-averaged transverse heat flux density at plate/fin surface (y = y0)as a function of the axial coordinate for two DR values.

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4527

The dependence of the cooling load _QC on the plate/fin spacing2y0 at selected LC values is shown in Fig. 11 (runs 18–23). At a fixedy0, the curve magnitude grows at increasing the fin length but thegrowth rate reduces with LC until the curves are being overlappingeach other for high enough LC values. This trend is consistent withthe results shown in Fig. 10 and should correspond to the flat por-tion of the curves attained for very long fins (LC� x1). Most impor-tant is that all curves peak at a well-defined value of the fininterspacing. The peak location results to be a slightly increasing(linear) function of the fin length, as evidenced in Fig. 11 by thestraight line passing through the locus of maxima. This should cor-respond to a reduction of the effective length available for heattransfer (i.e. of the HX length) for short plate spacing (y0/dj < 1).The optimum fin/plate spacing ranges between 2y0 = 2.98dj (forLC/2x1 = 0.079) and 2y0 = 3.34dj (for LC/2x1 = 1.9) for X = 0.76.These results agree with the general design procedure of selectingthe plate spacing of the stack of thermoacoustic refrigerators in theinterval 2dj 6 2y0 6 4dj [26].

The difference between the cooling load and the total axial en-thalpy flux evaluated at the midpoint of the stack (maximum cool-ing load) reveals that the HX performance is affected by some kindof thermal losses. To get insight into this issue the transverse com-ponent of the heat flux density at the gas-solid interface (y = y0) isplotted in Fig. 12 as a function of the axial coordinate (runs 24, 25).In both the upper (LC/2x1 = 0.53) and the lower (LC/2x1 = 2.25)curve illustrated the heat flux exhibits a sharp peak near the finends facing the duct (outer ends). These are therefore the mainareas where a net heat exchange between the fluid and the solidtakes place The lower plot evidences as the heat flux density in-verts its sign near the fin ends facing the stack (inner ends) of bothcold and hot HX giving rise to a reverse flux which decreases thenet heat transfer between the gas and the HXs, thus worseningthe HX performance. Smaller transverse heat fluxes, qualitativelysimilar to those observed in thermoacoustic couples [27], are alsofound near the plate ends but they quickly vanish in the middleplate regions. The extent of the reverse heat fluxes, which are qual-itatively similar to those found by Besnoin and Knio [13,14], re-duces with DR till to vanish at high drive ratios when LC/2x1 1,as evidenced by the upper curve of Fig. 12. An analogous trend isobserved if DR is held fixed and LC is progressively reduced. The re-verse flux effect or, in general, the poor efficiency of the fin areasfacing the stack (inner ends) in transferring heat could account

Fig. 11. The cooling load as a function of the plate/fin interspacing at selected LC

values and for DR = 4.93. Continuous lines result from the fit of the simulationpoints by an arbitrary peak function. The straight line passes through the locus ofmaxima.

for the previously discussed high performance of HXs whoselength is considerably smaller than 2x1. Reducing the length ofthe HX, in fact, results in decreasing the fin areas of poor efficiency(inner ends) so, even if the available heat transfer area becomessmaller, the heat transferred per unit area is increased. This isclearly observable in Fig 13 where the mean heat flux per unit sur-face area _Q �C=LC

exchanged by the cold fin with the gas is plotted

as a function of the fin length at selected drive ratios (runs 15–17).The area over which the HX is affected by the reverse flux also var-ies with the fin/plate spacing as illustrated in Fig. 14 (runs 18, 19,21) where the longitudinal profile of the transverse heat flux den-sity at the gas-solid interface (y = y0) of the cold HX fin is plotted atselected y0 values. The comparative analysis of this figure withFig. 11 informs that the points where _QC peaks are almost coinci-dent with the ones for which the reverse flux area results mini-mized. The lower plot of Fig. 12 reveals however that even whenthe reverse flux is eliminated, minor transverse heat fluxes local-ized at the stack edges persist.

The thermal losses above discussed affecting the HX-stack junc-tion seems therefore to constitute a permanent feature of the inter-element heat transfer process and cannot be totally avoided. Theresults of the present study informs however that their strengthcan be minimized by suitable choices of the fin length and spacingwhich thus have to be simultaneously optimized.

Fig. 13. The mean heat flux per unit surface area exchanged by the cold fin with thegas at selected drive ratios.

Page 11: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Fig. 14. The longitudinal profile of the time averaged transverse heat flux density atthe solid surface of the cold fin at selected fin spacing.

Fig. 15. The total cooling load per unit cross sectional area of the stack as a functionof blockage ratio at selected lengths of the cold HX.

4528 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

3.4. Effect of blockage ratio

The analysis performed in the last section indicates that theoptimum value of plate/fin spacing for which the cooling loadpeaks is around 3dj. This result was derived considering a constantblockage ratio, that is, in each run the plate thickness was ‘‘ad-justed’’ to give rise for the selected y0 to the constant valueX = 0.76. The analysis, therefore, did not take into account the ef-fect of varying plate/fin spacing on particle axial velocity and effec-tive cross sectional area available for heat transport. In order toestimate the effect of blocked ratio on the total cooling load simu-lations are performed at varying y0 for a fixed value of the platethickness l (roughly equal to the thermal penetration depth inthe fin material: 0.0005 m). Also, the total number of channelscomprised in a cross section of the stack are calculated assumingthat this last fill in all the resonator cross-section, taken for sim-plicity quadratic in shape and of unitary surface area. The totalcooling load sustained by the stack is then obtained by multiplying2 _QC for the resulting number of channels. Results are shown inFig. 15 where the total cooling load is reported as a function ofy0 (and X) at selected LC values (runs 26–28). All curves resultmuch more peaked compared to the ones reported in Fig. 11 andthe peak location (also in this case slightly dependent on LC) fallsat lower y0 values, i.e., at a fin/plate interspacing roughly equalto 2dj (to which the optimal value X � 0.59 corresponds). The dif-ferent behavior can be interpreted observing that the lower theplate interspacing the higher the number of channels of the stack.This compensates for the decreasing of the heat flux pumped in asingle channel which is observed in Fig. 11 to take place fory0 < 1.5dj. Simulations reveal that a change in plate thickness mod-ifies the peak height and associated optimum blockage ratio butnot the peak location. For example, doubling the plate thickness(for LC/2x1 = 1) reduces the peak height of about 9% and the optimalX by about 29% leaving the peak location unchanged (results notshown). Based on the results of this section the optimal plate/fininterspacing in stack-based thermoacoustic device should fall inthe narrower range 2dj 6 2y0 6 3dj.

3.5. Derivation of heat transfer correlations

In this section simulations are performed with the intent to de-velop heat transfer correlations which could be applied for the esti-mation of the gas-side heat transfer coefficient h at the HX walls. Atthe same time, the comparison of the theoretical predictions

against experimental data found in literature is considered asmodel validation.

Due to the existence of reverse heat flux localized at the inneredges of the HX fins the area-averaged h values over the cold HXare computed as indicated in [13] by the formula

h ¼ h _qyihTHXCi � hTgCi

ð31Þ

where

h _qyi ¼1LC

Z LC

0

_qyðx; y0Þdx ð32Þ

Analogous formulas have been applied to the hot HX. The h datathus obtained are converted to non-dimensional Nusselt numbersbased on the hydraulic diameter of the pore Dh (=4y0) and then toColburn-j factors

NuD ¼hDh

K; jC;D ¼

NuD

Re1;DPr1=3 ð33Þ

Re1,D [=vx1(0)Dh/m] being the acoustic Reynolds number defined interms of Dh and of the amplitude of the acoustic velocity evaluatedat gas centerline. The results of about 40 simulations (runs 29–37)carried out at varying DR, LC, LH, y0, and xS values are illustrated inFig. 16 where the resulting Colburn-j factors are plotted as a func-tion of Re1,D. In the same graph the experimental measurementsof Nsofor et al. [28], Peak et al. [29], Mozurkewich [9] and Brewsteret al. [22] are reported. All HXs involved in these works are of thefinned-tube type (except the one in [9] consisting of a copper screensoldered to tubes referred to by the author as ‘‘configuration C’’) sothe measured h values therein reported are considered suitable tomake comparison with the predictions of the proposed model. Pre-liminary, in order to make the different set of data consistent, theexperimental Nusselt and Reynolds numbers have been re-scaledby the same length parameter (4y0). In addition, the rms velocityamplitudes entering in the Reynolds numbers reported in [5] havebeen ‘‘corrected’’ by the factor

ffiffiffi2p

and by the blockage ratio X.The mean deviation of the predicted jC values from the experi-

mental ones amounts to 49% for the data of Peak et al., 36% for thedata o Brewster et al., 36% for the data of Mozurkewich and 56%for the data of Nsofor et al. Deviations in the moderate and high Rey-nolds number regime (Re1 > 500) are expected because of the in-creased weight of entrance/exit effects at the resonator-HX cross

Page 12: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

Fig. 16. Comparison between experimental and theoretical Colburn-j factors.

A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530 4529

section interface. These effects are responsible of complex non-lin-ear temperature and flow patterns (turbulent and vorticity flows[6].) which highly affect heat transfer rates causing, among otherthings, minor losses and heat leakage to the duct. The high deviationfound at low Re1 numbers (Re1 < 500) between simulations andexperimental results of Nsofor et al. is quite surprising becausethe linear theory, on which the present model relies upon, is gener-ally observed to match experiments quite well at low velocityamplitudes. In the evaluation of these results it has to be pointedout however that, although proper scaling procedures have been ap-plied, the operating and geometrical conditions specified in the sim-ulations are different from those of the selected experiments. It’sremarkable to note, anyway, that al low Re1 values (Re1 < 200) thesimulation points match exactly the correlation curve derived byPeak et al. from the fit of their data. On the same curve, on the otherhand, the data of Mozurkewich tend to collapse at decreasing Re1.

4. Conclusions

The performance of parallel-fin thermoacoustic HXs coupled toa stack in a simplified configuration is investigated through a 2Dnumerical model based on the classical linear thermoacoustic the-ory. Details of the transverse heat transfer at the HXs are high-lighted and optimization information for their design are derived.The following conclusion can be drawn:

1. The magnitude of the heat transfer coefficient from the second-ary fluid-side (U) should be fully taken into account because itaffects cooling load through its influence on the temperaturedistribution across the stack. High U values are desirable toimprove the performance of the device.

2. Fin length along the axial direction of particle oscillation can bechosen considerably lower than the peak-to-peak displacementamplitude without compromising the HX performance andwith great benefits in terms of viscous losses reduction.

3. Thermal losses (reverse heat fluxes) localized at the stack-HXjunctions degrade the HXs performance reducing the usefulcooling load. These effects could account for about 7% of thedeviations found between predictions of the linear theory andexperimental measurements. Suitable choices of the fin spacingand fin length can contribute to minimize this detriment effectsenhancing the effectiveness of the HXs.

4. Optimum fin/plate interspacing should fall in the range2dj 6 2y0 6 4dj. Selection of the best value is case dependentand should also take into account the simultaneous effect ofblockage ratio, fin length and fin thickness.

5. Heat transfer coefficients from the gas-side can be predictedwith a confidence of 36% and 49% respectively at moderateand high acoustic Reynolds numbers. The high deviation foundat low Re1 with available experimental data requires furtherinvestigation.

The simplified computational method proposed in this worksuffers by several limitations. In its current form the model is re-stricted to the acoustic and short stack approximations. Further-more, the HXs are modeled as coupled to the stack in asimplified configuration while their arrangement in real TA devicescan be more complex. The heat transport equations implementedin the code relies on the classical form of the function T1 whoseaccuracy is questionable near the pore terminations. The modeldoes not take into account for entrance and exit effects whichmay become relevant near sudden changes of cross section likethose existing at the ends of the HXs. Further analysis is needed,particularly of the fluid and energy flows near the HX pore ends,for overcoming these shortcomings.

Acknowledgments

This research is financially supported by the EC under the pro-ject framed within the SEVENTH FRAMEWORK PROGRAMME‘‘THermoacoustic Technology for Energy Applications’’ (GrantAgreement No.: 226415, Thematic Priority: FP7-ENERGY-2008-FET, Acronym: THATEA). The author also gratefully acknowledgesthe allocation of computer time from the ‘‘Centro di Calcolo Elet-tronico Attilio Villari’’ of the University of Messina.

References

[1] G.W. Swift, Thermoacoustic engines, J. Acoust. Soc. Am. 84 (1988) 1145–1180.[2] G.W. Swift, Thermoacoustics: a unifying perspective for some engines and

refrigerators, Acoust. Soc. Am., 2002.[3] S. Backhaus, G.W. Swift, A thermoacoustic Stirling heat engine: detailed study,

J. Acoust. Soc. Am. 107 (2000) 3148–3166.[4] K. Tang, R. Bao, G.B. Chen, Y. Qiu, L. Shou, Z.J. Huong, T. Jin, Thermoacoustically

driven pulse tube cooler below 60 K, Cryogenics 47 (9–10) (2007) 526–529.[5] M.E.H. Tijani, S. Spoelstra, Study of coaxial thermoacoustic Stirling cooler,

Cryogenics 48 (2008) 77–82.[6] L. Shi, Z.B. Yu, A.J. Jaworski, Vortex shedding flow patterns and their transitions

in oscillatory flows past parallel-plate thermoacoustic stacks, Exp. ThermalFluid Sci. 34 (7) (2010) 954–965.

[7] A. Berson, M. Michard, P. Blanc-Benon, Measurement of acoustic velocity in thestack of a thermoacoustic refrigerator using particle image velocimetry, HeatMass Transfer 44 (8) (2008) 1015–1023.

[8] A. Piccolo, G. Pistone, Estimation of heat transfer coefficients in oscillatingflows: the thermoacoustic case, Int. J. Heat Mass Transfer 49 (9–10) (2006)1631–1642.

[9] G. Mozurkewich, Heat transfer from transverse tubes adjacent to athermoacoustic stack, J. Acoust. Soc. Am. 110 (2) (2001) 841–847.

[10] V. Gusev, P. Lotton, H. Bailliet, S. Job, M. Bruneau, Relaxation-timeapproximation for analytical evaluation of temperature field inthermoacoustic stack, J. Sound Vib. 235 (5) (2000) 711–726.

[11] G. Mozurkewich, A model for transverse heat transfer in themoacoustics, J.Acoust. Soc. Am. 103 (1998) 3318–3326.

[12] A.S. Worlikar, O.M. Knio, Numerical study of oscillatory flow and heat transferin a loaded thermoacoustic stack, Numer. Heat Transfer Part A 35 (1999) 49–65.

[13] E. Besnoin, O. M Knio, Numerical study of thermoacoustic heat exchangers inthe thin plate limit, Numer. Heat Transfer A 40 (2001) 445–471.

[14] E. Besnoin, O. M Knio, Numerical study of thermoacoustic heat exchangers,Acta Acust. United Acust. 90 (2004) 432–444.

[15] D. Marx, P. Blanc-Bennon, Numerical simulation of stack–heat exchangerscoupling in a thermoacoustic refrigerator, AIAA J. 42 (2004) 1338–1347.

[16] A. Piccolo, G. Pistone, Computation of the time-averaged temperature fieldsand energy fluxes in a thermally isolated thermoacoustic stack at low acousticMach numbers, Int. J. Thermal Sci. 46 (3) (2007) 235–244.

[17] L.D. Landau, E.M. Lifshitz, Fluid Mechanics, first ed., Pergamon, London, 1959.[18] R.S. Wakeland, R.M. Keolian, Effectiveness of parallel-plate heat exchangers in

thermoacoustic devices, J. Acoust. Soc. Am. 115 (6) (2004) 2873–2886.[19] W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannerry, Numerical Recipes

in Fortran, Snd ed., Cambridge University Press, Berlin, 1994.[20] http://www.netlib.org/lapack.

Page 13: Numerical computation for parallel plate thermoacoustic heat exchangers in standing wave oscillatory flow

4530 A. Piccolo / International Journal of Heat and Mass Transfer 54 (2011) 4518–4530

[21] P. Merkli, H. Thomann, Transition to turbulence in oscillating pipe, J. FluidMech. 68 (1975) 567–579.

[22] J.R. Brewster, R. Raspet, H.E. Bass, Temperature discontinuities betweenelements of thermoacoustic devices, J. Acoust. Soc. Am. 102 (6) (1997).

[23] G.W. Swift, Analysis and performance of a large thermoacoustic engine, J.Acoust. Soc. Am. 92 (1992) 1551–1563.

[24] D. Marx, P. Blanc-Benon, Computation of the temperature distortion in thestack of a standing-wave thermoacoustic refrigerator, J. Acoust. Soc. Am. 118(5) (2005) 2993–2999.

[25] T.J. Hofler, Effective heat transfer between a thermoacoustic heat exchangerand stack, J. Acoust. Soc. Am. 94 (1993) 1772.

[26] M.E.H. Tijani, J.C.H. Zeegers, A.T.A.M. de Waele, The optimal stack spacing forthermoacoustic refrigeration, J. Acoust. Soc. Am. 112 (1) (2002) 128–133.

[27] L. Zoontjens, C.Q. Howard, A.C. Zander, B.S. Cazzolato, Numerical study of flowand energy fields in thermoacoustic couples of non-zero thickness, Int. J.Thermal Sci. 48 (4) (2009) 733–746.

[28] E.C. Nsofor, S. Celik, X. Wang, Experimental study on the heat transfer at theheat exchanger of the thermoacoustic refrigerating system, Appl. Thermal Eng.27 (2007) 2435–2442.

[29] I. Paek, J.E. Braun, L. Mongeau, Characterizing heat transfer coefficients for heatexchangers in standing wave thermoacoustic coolers, J. Acoust. Soc. Am. 118(4) (2005) 2271–2280.