babaei design and optimization of thermoacoustic devices

Upload: don-degrasse

Post on 08-Jul-2018

221 views

Category:

Documents


1 download

TRANSCRIPT

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    1/14

    Design and optimization of thermoacoustic devices

    Hadi Babaei, Kamran Siddiqui *

    Department of Mechanical and Industrial Engineering, Concordia University, Montreal, Canada

    a r t i c l e i n f o

     Article history:

    Received 6 December 2007

    Accepted 21 July 2008

    Available online 11 September 2008

    Keywords:

    Thermoacoustics

    Sustainable refrigerator

    Design procedure

    a b s t r a c t

    Thermoacoustics deals with the conversion of heat energy into sound energy and vice versa. It is a new

    and emerging technology which has a strong potential towards the development of sustainable and

    renewable energy systems by utilizing waste heat or solar energy. Although simple to fabricate, the

    designing of thermoacoustic devices is very challenging. In the present study, a comprehensive design

    and optimization algorithm is developed for designing thermoacoustic devices. The unique feature of 

    the present algorithm is its ability to design thermoacoustically-driven thermoacoustic refrigerators that

    can serve as sustainable refrigeration systems. In addition, new features based on the energy balance are

    also included to design individual thermoacoustic engines and acoustically-driven thermoacoustic refrig-

    erators. As a case study, a thermoacoustically-driven thermoacoustic refrigerator has been designed and

    optimized based on the developed algorithm. The results from the algorithm are in good agreement with

    that obtained from the computer code DeltaE.

     2008 Elsevier Ltd. All rights reserved.

    1. Introduction

    Thermoacoustic is a branch of science dealing with the conver-

    sion of heat energy into sound energy and vice versa. Device that

    converts heat energy in sound or acoustic work is called thermoa-

    coustic heat engine or prime mover and the device that transfers

    heat from a low temperature reservoir to a high temperature res-

    ervoir by utilizing sound or acoustic work is called thermoacoustic

    refrigerator. Although the thermoacoustic phenomenon was dis-

    covered more than a century ago, the rapid advancement in this

    field occurred during the past three decades when the theoretical

    understanding of the phenomenon was developed along with the

    prototype devices based on this technology [1,2]. The thermoacou-

    stic technology has not reached the technical maturity yet, as a re-

    sult, the performance of thermoacoustic devices is still lower than

    their convectional counterparts. Thus, significant efforts are

    needed to bring this technology to maturity and develop compet-

    itive thermoacoustic devices. There are several advantages of heat

    engines and refrigerators based on thermoacoustic technology as

    compared to the conventional ones. These devices have fewer com-

    ponents with at most one moving component with no sliding seals

    and no harmful refrigerants or chemicals are required. Air or any

    inert gas can be used as working fluids which are environmentally

    friendly. Furthermore, the fabrication and maintenance costs are

    low due to inherent simplicity of the thermoacoustic devices.

    The main components of a typical thermoacoustic engine or

    refrigerator are a resonator, a stack of parallel plates and two heat

    exchangers. A half wavelength (or a quarter wavelength) acoustic

    standing wave is generated in the resonator. The thermoacoustic

    phenomenon takes place in the stack when a nonzero temperature

    gradient imposed along the stack plates (i.e. parallel to the direc-

    tion of the sound wave propagation) interacts with the sound wave

    oscillations. The heat exchangers are responsible of transferring

    heat in and out of a thermoacoustic device at their desired temper-

    atures, thus maintaining a given temperature gradient along the

    stack.

    Thermoacoustic refrigerators can be classified based on the

    source of the acoustic energy input. If the acoustic energy is pro-

    vided by a thermoacoustic engine, the refrigerator is called ther-

    moacoustically-driven thermoacoustic refrigerator (TADTAR).

    Whereas, if the acoustic energy is provided by an acoustic driver

    e.g. a loudspeaker, it is termed as acoustically-driven thermoacou-

    stic refrigerator. During the past decades, several acoustically-dri-

    ven thermoacoustic refrigerators have been developed   [3–5].

    Although the form of energy consumed in these refrigerators is

    acoustic, the energy source for the acoustic driver is typically elec-

    trical from conventional energy resources. During recent years,

    there is an increased interest in the development of thermoacous-

    tically-driven thermoacoustic refrigerators. These devices are built

    by coupling a thermoacoustic refrigerator to a thermoacoustic en-

    gine. Thermoacoustic engines are capable of producing acoustic

    energy from any source of heat energy. Thus, the primary energy

    source to drive the refrigerator could be conventional or unconven-

    tional that includes industrial waste heat, solar energy and fossil

    fuels. If the heat source for the thermoacoustic engine is the indus-

    trial waste heat or solar energy then this device has two major

    advantages. Firstly, it does not require any addition conventional

    0196-8904/$ - see front matter   2008 Elsevier Ltd. All rights reserved.doi:10.1016/j.enconman.2008.07.002

    *  Corresponding author. Tel.: +1 514 848 2424x7940; fax: +1 514 848 3175.

    E-mail address: [email protected] (K. Siddiqui).

    Energy Conversion and Management 49 (2008) 3585–3598

    Contents lists available at   ScienceDirect

    Energy Conversion and Management

    j o u r n a l h o m e p a g e :   w w w . e l s e v i e r . c o m / l o c a t e / e n c o n m a n

    mailto:[email protected]://www.sciencedirect.com/science/journal/01968904http://www.elsevier.com/locate/enconmanhttp://www.elsevier.com/locate/enconmanhttp://www.sciencedirect.com/science/journal/01968904mailto:[email protected]

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    2/14

    energy resource and secondly, by utilizing the waste heat, theamount of total waste heat rejected to the thermal energy sink will

    be reduced which will increase the overall performance of the en-

    tire system. Thus, a complete thermoacoustic refrigeration system

    in which the heat engine (which operates on waste heat) drives a

    refrigerator and the entire system has no harmful affects on the

    environment can be termed as a ‘‘sustainable refrigeration sys-

    tem”. In contrast to the acoustically-driven thermoacoustic refrig-

    erator which has one moving component i.e. the acoustic driver,

    thermoacoustically-driven thermoacoustic refrigerator has no

    moving parts thus; chances of mechanical failure are extremely

    low.

    Recently, some efforts have been made to develop heat engines

    that operate on waste heat. Symko et al. [6] designed and devel-

    oped a thermoacoustic heat engine that utilizes heat from a micro-circuit to produce sound. Hatazawa et al.  [7]   proposed a heat

    engine that utilizes waste heat from a four-stroke automobile gas-

    oline engine. Adeff and Hofler [8] developed a prototype thermoa-

    coustic refrigeration system that operates on the solar energy.

    Babaei et al. [9] have proposed a thermoacoustic refrigeration sys-

    tem for a gas turbine trigeneration system that operates on the

    waste heat from the gas turbine. It has been demonstrated that

    the thermoacoustic refrigeration system has the ability to enhance

    the overall efficiency of a trigeneration system by 5%.

    Some recent theoretical studies have demonstrated the strong

    potential of thermoacoustic devices in energy conservation and

    reduction of harmful emissions. A study shows that if all the indus-

    trial waste heat above 140 C in Netherlands can be used in ther-

    moacoustic devices, this would save 16 PJ per year whichcorresponds to the saving of more than 5 billion m3 of natural

    gas  [10]. It is estimated that over 32 billion liters of fuel is con-sumed annually for the operation of vehicle air-conditioners in

    the US alone. Modern vehicle refrigeration systems use R-134a,

    with a global warming potential still 1300 times that of carbon

    dioxide   [11]. Zoontjens et al.   [12]   theoretically investigated the

    feasibility of using thermoacoustic devices as the air conditioning

    system of an automotive by utilizing the automotive waste heat.

    They concluded that the thermoacoustic refrigerator has a strong

    potential to replace the existing automotive air conditioning

    systems.

    Although thermoacoustic devices are easy to build and main-

    tain, designing of these devices involves significant technical chal-

    lenges. These challenges become more substantial when designing

    a thermoacoustically-driven thermoacoustic refrigerator. This is

    attributed to the complicated thermoacoustic theory which is notdirectly applicable for design purposes. Thus, a systematic ap-

    proach is necessary to design and optimize thermoacoustic

    devices.

    Wetzel and Herman   [13]   developed a design algorithm for

    acoustically-driven thermoacoustic refrigerators. They developed

    the design algorithm by using the simplified linear thermoacoustic

    model, and normalizing the position and length of the refrigerator

    stack and the equations of the total power flow and consumed

    acoustic power in the stack. By applying the algorithm, the de-

    signer can decide the stack position and length at the given tem-

    peratures of heat exchangers to have the maximum performance

    of the stack. The geometrical parameters such as stack plate thick-

    ness and spacing as well as the cross-sectional area of the resona-

    tor can also be estimated. In this study, however, it is not describedhow the desired cooling power of the refrigerator, the stack

    Nomenclature

     A   cross-sectional area (m2)a   speed of sound (m/s)BR blockage ratioCOP coefficient of performanceCOPR coefficient of performance relative to Carnot

    c  p   isobaric heat capacity of the working gas (J kg1 K1)c solid   heat capacity of the stack plates (J kg

    1 K1)DR drive ratio_E 2   work flux (W)D _E 2   produced or consumed work flux (W) f    resonant frequency (Hz)_H 2   total energy flux (W)

    HX heat exchangerK    thermal conductivity of the working gas (W m1 K1)k   wave number (m1)L   length (m)l   half of the plate thickness (m)P m   mean pressure (Pa)P A   antinode pressure amplitude (Pa) p1   pressure amplitude (Pa)

    Pr    Prandtl numberQ    heat flux (W)r h   hydraulic radius (m)S    surface area (m2)_S    entropy flux (W/K)T    temperature (K)DT    temperature difference (K)rT    temperature gradient (K/m)U 1   volume flow rate (m

    3/s)u1   velocity amplitude (m/s) x1   gas displacement amplitude (m) xc   stack center position (m)

     y0   half of the plate spacing (m)a   thermal diffusivity (m2 s1)b   thermal expansion coefficient (K1)dk   thermal penetration depth (m)dv   viscous penetration depth (m)

    es   plate heat capacity ratioc   ratio, isobaric to isochoric specific heatsC   normalized temperature gradientgth   thermal efficiencyk   wavelength (m)l   dynamic viscosity (kg m1 s1)P   perimeter (m)q   density (kg m3)x   angular frequency (rad s1)

    Subscripts, superscriptsa ambientc coldcon consumedcrit criticald ducteng enginegen generatedh hotm meann normalizedpro producedr resonatorref refrigerators stackt total

    3586   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    3/14

    consumed acoustic power and the total power flow in the stack are

    correlated, and under which configuration of the refrigerator stack

    this correlation is valid.

    Tijani et al. [14] also described the design algorithm for acous-

    tically-driven thermoacoustic refrigerators by considering a corre-

    lation between the desired cooling power of the refrigerator, the

    stack consumed acoustic power and total power in the stack,

    which is different from that of Wetzel and Herman [13]. It is how-ever, not well described how this correlation is derived and at

    which refrigerator configuration it may be applied.

    The above design algorithms are applicable only to design

    acoustically-driven thermoacoustic refrigerators. These algorithms

    cannot be used in designing a thermoacoustically-driven thermoa-

    coustic refrigerator (TADTAR), as designing of TADTAR involves

    more parameters and it is more challenging than the acousti-

    cally-driven thermoacoustic refrigerators. Therefore, to design

    and develop efficient sustainable thermoacoustic refrigeration sys-

    tems, a detailed design and optimization procedure is necessary.

    To the best of authors’ knowledge no such design and optimization

    procedure or algorithm is available.

    In this paper, a comprehensive systematic procedure has been

    developed for the design and optimization of thermoacoustic de-

    vices by applying the simplified linear thermoacoustic model. This

    procedure which is mainly intended to design and optimize a ther-

    moacoustically-driven thermoacoustic refrigerator (TADTAR) can

    also be used to design and optimize individual thermoacoustic en-

    gines and acoustically-driven thermoacoustic refrigerators. It

    should be noted that the procedure presented in this study pro-

    vides a more comprehensive discussion on the design and optimi-

    zation of acoustically-driven thermoacoustic refrigerators than

    previous studies. The design procedure which is based on the en-

    ergy and entropy balances applied on different components of 

    the device is a simple and effective tool to design and optimize a

    thermoacoustic device to meet its requirements. The goal of 

    designing a thermoacoustically-driven thermoacoustic refrigerator

    is to meet the required cooling power at the desired cooling tem-

    perature and at the given heat input temperature while rejectingsome heat to the environment.

    The developed algorithm not only provides a step by step pro-

    cedure to design and optimize a thermoacoustic device but also en-

    ables to evaluate the influence of different parameters on the

    behavior and performance of the device.

    Finally, a thermoacoustically-driven thermoacoustic refrigera-

    tor is designed and optimized based on the developed procedure

    and simulated by the computer code DeltaE to compare and verify

    the design parameters.

    It is worth mentioning that using DeltaE to design thermoacou-

    stic devices from scratch needs tremendous amount of effort espe-

    cially in the case of thermoacoustically-driven thermoacoustic

    refrigerator. The presented procedure significantly reduces the

    technical challenges associated with the designing of thermoacou-stic devices.

    2. Thermoacoustic principle

    Phasing plays an important role in the operation of thermoa-

    coustic devices. To attain a proper phasing in a thermoacoustic de-

    vice, a rather poor thermal contact is essential between the gas

    parcel and its adjacent solid plate. This imperfect thermal contact

    causes the heat flow between the gas and the plate, not to produce

    instantaneous changes in the gas temperature. Instead, the heat

    flow creates a time phasing between temperature, pressure and

    displacement needed to drive the gas parcels through a thermody-

    namic cycle [2].

    Consider a solid plate aligned in the direction of the acousticwave propagation with an imposed temperature gradientrT  along

    the plate. The length of the plate is assumed equal to the peak to

    peak displacement of gas parcels (2 * j x1j) oscillating along the

    plate (see Fig. 1a). The figure shows a magnified view of a single

    plate stack and a gas parcel oscillating next to it in a half wave-

    length thermoacoustic refrigerator. The gas parcel oscillates under

    the influence of standing wave generated by the acoustic power in-

    put. The heat energy is transferred in and out of the device by the

    cold and ambient heat exchangers located at the edges of the stackplate, respectively. The temperatures of the heat exchangers im-

    pose a temperature gradient along the stack plate (rT ). The varia-

    tion of the pressure and velocity magnitudes of the acoustic wave

    along the resonator is shown in Fig. 1b.

    In a real thermoacoustic device, the oscillations are sinusoidal;

    but for simplicity, the square wave motion is considered to explain

    the basic thermodynamic cycle that a gas parcel undergoes. The

    gas parcel experiences two adiabatic processes while moving along

    the solid plate and two irreversible constant pressure processes

    while exchanging heat with the solid plate [2]. Two temperatures

    are important to the parcel. The temperature of the gas parcel after

    adiabatic compression and expansion (imposed by the sound wave

    and related to the sound wave pressure oscillation) and the local

    temperature of the solid plate (imposed by the heat exchangers)

    adjacent to the gas parcel after adiabatic compression and expan-

    sion and displacement of the gas parcel. Note that the acoustic

    wave is responsible for both adiabatic compression and expansion,

    and the displacement of the gas parcel along the solid plate. If the

    temperature of the gas is higher than that of the plate, heat flows

    from the gas to the plate. If the temperature of the gas is lower

    than that of the plate, heat flows from the plate to the gas. Thus,

    it is the imposed temperature gradient  rT   along the plate that

    makes a thermoacoustic device to operate as an engine or a refrig-

    erator. A zero or low temperature gradient is the condition for a

    refrigerator and a high temperature gradient is the condition for

    an engine. If  rT   along the plate be selected in such a way that

    the temperature change along the plate (2rT j x1j) as seen by the

    parcel just matches the parcel’s temperature change due to adia-

    batic compression and expansion 2 T mb p1qmc  p

    , no heat would flow

    between the parcel and the solid plate. This temperature gradient

    is called the critical temperature gradient and is defined as  [2],

    Fig. 1.  (a) Schematic of a thermoacoustic refrigerator with a single plate stack, (b)

    variation of pressure and velocity amplitudes along the resonator tube, solid line:pressure, dashed line: velocity.

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3587

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    4/14

    rT crit ¼ T mbx p1qmc  pu1ð1Þ

    Usually the length of the plate is larger than the displacement of a

    given gas parcel. Thus, the heat transfer from one end of the plate to

    the other end is occurred by a series of gas parcels along the plate as

    depicted in Fig. 2. In  Fig. 2, gas parcel A  absorbs heat from the plate

    at location  a  and transfers it to the plate at location  b. Half a cycle

    later, the adjacent gas parcel  B  picks this heat from the location  b

    (at this instant, parcel A  is at the location  a). This heat is transferred

    to location c  by the parcel B from where the gas parcel C  picks it and

    delivers to the location d. Thus, the heat is transferred from one end

    to the other by gas parcels as bucket brigade. It should be noted that

    the plate is used only for the temporary storage of heat [2].

     2.1. Simplified linear model of thermoacoustic devices

    Consider a stack of parallel plates with x axis along the direction

    of the acoustic wave propagation and  y   axis perpendicular to the

    plane of the stack. The plate thickness is equal to 2l and the plate

    spacing is equal to 2 y0. The simplified thermoacoustic model is

    developed by linearizing momentum, continuity and heat flow

    equations, and considering the following three assumptions [2].First, it is assumed that y0dj,  y0  dv  where  dk   is the thermal

    penetration depth defined as the thickness of the layer around

    the stack plate through which the heat can diffuse in the fluid,

    whereas,   dm   is the thickness of the layer around the stack plate

    where the viscous effects are significant. This assumption is called

    the boundary layer approximation and typically in thermoacoustic

    devices, dj 6 y0 6 2dj [2]. Second, the length of the stack is consid-

    ered to be significantly less than the wavelength of the standing

    acoustic wave (i.e. Ls k) such that it does not perturb the acoustic

    standing wave (short stack approximation). With this approxima-

    tion and assuming standing wave phasing between pressure and

    velocity, the velocity and pressure can be expressed as [2],

     p1 ¼ P A cosðkxÞ;   u1 ¼   1 þ   l y0

      P Aqma

    sinðkxÞ ð2Þ

    Finally, it is assumed that the stack is short enough that  p1  and  u1could be regarded as independent of   x   within the stack, and the

    temperature difference along the stack is less than the stack mean

    temperature (i.e.  DT T m). So the thermophysical properties of the

    gas are assumed to be independent of  x  within the stack. Thus,  p1,

    u1  and thermophysical properties are evaluated at the stack mid

    point, i.e. the stack mid temperature [2].

    The simple linear expressions for total power flow   _H 2  (i.e. total

    energy flux) through the stack, and the acoustic power D _E 2   (work

    flux) produced in (or consumed by) the stack are expressed by

    the following equations [2,15]:

    _H 2 ¼ A4

    dk

    r h

    T mb j p1kU 1 j Að1þ esÞð1þPr ÞK   C

    1þ ffiffiffiffiffiPr 

    p   þPr þPr es

    1þ ffiffiffiffiffiPr 

    p      1þ ffiffiffiffiffiPr 

    p   dv

     y0

    " #

    ð AK þ AsolidK solidÞrT    ð3ÞD _E 2 ¼ A

    4

    Lsr h

    ðc1Þ j p1j2dkxc pmð1þesÞ

    C

    1þ ffiffiffiffiffiPr 

    p  K

    10@

    1Aqm jU 1j2dvx

     A2K

    24

    35

    ð4Þwhere  C   is the normalized temperature gradient defined as   rT 

    rT crit,

    K ¼ 1 dvr hþ

      d2v2r 2

    h

    ,   A  and   Asolid  are the fluid and solid cross-sectional

    areas in the stack, respectively, andes is the plate heat capacity ratio

    defined as,  es  ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 

    qmc  p K qsolidc solidK solid

    q   .

    By assuming that all dimensions of the resonator are much lar-

    ger than the penetration depths and the temperature gradient

    along the axis of the resonator is zero, the acoustic power loss

    per unit surface area of the resonator can be estimated as [15],

    d _E 2dS 

     ¼ 14qm

    U 1 A

    2

    dvx 14

    j  p1j2c pm

    ðc 1Þdjx   ð5Þ

    The first term on the right hand side of Eq.  (5) represents the energy

    dissipated due to the viscous shear and the second term on the right

    hand side represents the energy dissipated due to the thermal

    relaxation.

    It should be noted that the stack plates are assumed ideal so

    that the plate heat capacity ratio  es   is zero and the last term inthe right hand side of Eq. (3) which represents the axial conduction

    in the stack plates and the working gas is neglected. These two

    terms have a negligible effect on the calculations [13]. By consider-

    ing only ideal gases close to their critical point as the working gas,

    the parameter T mb  in Eq. (3)  can be set equal to unity [2].

    3. Design and optimization procedure

    Besides the available features from previous studies, following

    new features are applied in the present study to develop the com-

    prehensive design and optimization procedure for thermoacoustic

    devices.

    The simplified linear thermoacoustic model is used to evaluate

    the engine part of the device. All the dimensions in the direction of 

    the acoustic wave propagation including the length and position of 

    the stacks and heat exchangers are normalized. The normalized

    acoustic power equation is applied to estimate the dissipated

    acoustic power in the heat exchangers. The equation estimatingthe acoustic power losses in the resonator’s wall surface area is

    normalized (see Eq. (9)). A comprehensive discussion is presented

    to correlate the desired cooling power and the required heat input

    to the total power flow in the stack and the acoustic power flow, by

    applying the energy balance on the cold and hot heat exchangers

    (see Eqs.   (11), (15), (16), (18), and (20)). The normalized engine

    stack position and length are selected by applying the energy bal-

    ance on the whole device and these selections are then modified to

    have the engine stack performs at the maximum efficiency at the

    given temperatures of the heat exchangers. This behavior is also

    shown by applying the entropy balance and energy balance on a

    device, simultaneously. It is shown that the engine stack position

    and length could be selected to have the minimum entropy gener-

    ation within the system while producing the required acousticpower to run the system at the desired temperatures of the heatFig. 2.  Mechanism of heat transfer by the gas parcels along the stack plate of athermoacoustic device.

    3588   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    5/14

    exchangers (Eq.   (33)). The designer could also estimate the re-

    quired heat input at the desired temperature to run the engine sec-

    tion of the device (Eq. (35)).

    Heat exchangers are the least understood components of ther-

    moacoustic devices and their proper designing is a critical task,

    as the literature contains very little experimental or analytical

    guidance. Swift   [2] originally argued that the optimum length of 

    a heat exchanger should be equal to the peak to peak gas displace-ment amplitude at the heat exchanger location. The ambient heat

    exchanger is always closer to the velocity node, so the peak to peak

    gas displacement at its location is smaller than that of the cold heat

    exchanger in a thermoacoustic refrigerator. On the other hand, the

    ambient heat exchanger transfers more heat compared to the cold

    heat exchanger in a thermoacoustic refrigerator as it must transfer

    both the transferred heat by the cold heat exchanger and portions

    of dissipated acoustic power in the device to the outside environ-

    ment. So it is safe to say that by assuming the same heat transfer

    coefficient and temperature difference between the solid plate

    and the working gas, the ambient heat exchanger requires more

    heat transfer area compared to the cold heat exchanger. The same

    argument can be raised for thermoacoustic engines. The heat

    transferred by the ambient heat exchanger in a thermoacoustic en-

    gine is smaller than that transferred by the hot heat exchanger

    while the hot heat exchanger is always closer to the velocity node.

    For the design procedure, following assumptions are made for

    heat exchangers. All heat exchangers are assumed parallel plate.

    The length of the cold heat exchanger and ambient heat exchanger

    in a thermoacoustic refrigerator are assumed equal to the peak-to-

    peak gas displacement amplitude at the cold heat exchanger loca-

    tion. The length of the hot heat exchanger and ambient heat

    exchanger in a thermoacoustic engine are assumed equal to the

    peak-to-peak gas displacement amplitude at the ambient heat ex-

    changer location. The blockage ratio of heat exchangers is assumed

    equal to that of their respective stack.

    In the following subsections, normalization of thermoacoustic

    parameters and equations are described first followed by the

    description of the energy balance (first law of thermodynamics)and entropy balance (second law of thermodynamics) on the se-

    lected control volumes of the device.

     3.1. Normalization

    The energy flux equations (Eqs. (3)–(5)) indicate that there are

    different sets of independent parameters that play important roles

    in evaluating the performance of a thermoacoustic system. These

    parameters can be categorized in three main categories. Geometri-

    cal variables, which are stack plate thickness and spacing, position

    and length of the stack, and stack cross-sectional area. Material re-

    lated variables that include thermophysical properties of the work-

    ing gas and the stack. Design related variables which are resonance

    frequency, mean pressure and pressure amplitude of the workinggas, mean temperature and temperature difference along the stack

    and the desired cooling power of the system  [13]. Due to a large

    number of design parameters, the number of independent param-

    eters can be reduced through normalization.  Table 1   shows the

    independent variables, normalizing parameters and the normal-

    ized form [16]. In the present study, the maximum value for the

    normalized stack position (measured from pressure antinode)

    and the normalized stack length are assumed equal to 0.5 to avoid

    large viscous dissipation which decreases the overall performance

    of the device.

    As recommended in the literature, the normalized plate spacing(i.e. blockage ratio) was set equal to 0.8 [2,4,17].

    The normalized thermal penetration depth and the normalized

    viscous penetration depth can be assumed in the range 0.5–1 and

    0.5 Pr 2 to Pr 2, respectively [2].

    As the thermoacoustic model is based on the linear wave the-

    ory, to avoid nonlinearities, it is recommended that the drive ratio

    (DR = p1/ pm) should be smaller than 3% so that the acoustic Mach

    number and acoustic Reynolds number would be smaller than

    0.1 and 500, respectively [18,19].

    The normalized temperature difference along the refrigerator

    stack and engine stack are assumed in the range 0–0.17 and

    0.35–0.95, respectively. The mean temperature along the refriger-

    ator stack and engine stack are assumed in the range 288–303 K,

    and 390–600 K, respectively. It should be noted that to satisfy

    the third assumption described in the previous section, thermo-

    physical properties of the gas inside the refrigerator stack (and res-

    onator) are calculated based on the refrigerator stack mean

    temperature and the thermophysical properties of the gas inside

    the engine stack are calculated based on the engine stack mean

    temperature.

    The normalized temperature gradient, C   ¼ DT =LsrT crit

    ¼   rT rT crit

     can be

    expressed as the function of other normalized parameters as [2],

    C ¼   DT nBR ðc 1ÞLsn cotð xcnÞ   ð6Þ

    The above equation shows that for a stack with specified length and

    position, there is a range of normalized temperature differences at

    which the stack operates as a refrigerator (C

    < 1) and there is arange of normalized temperature differences at which it operates

    as an engine (C > 1).

    By dividing the total power and acoustic power equations (Eqs.

    (3) and (4)) by the product   AP ma, and assuming a parallel plate

    stack (r h =  y0), the following normalized equations for the total

    power flow through the stack   ð _H 2nÞ  and the acoustic power pro-

    duced in (or consumed by) the stack ðD _E 2nÞ are expressed as [16],

    _H 2n    18c

    dknDR 2  sinð2 xcnÞð1 þ Pr ÞK   C

    1 þ ffiffiffiffiffiPr 

    p   þ Pr 

    1 þ ffiffiffiffiffiPr 

    p      1 þ ffiffiffiffiffiPr 

    p   dvn

    " #

    ð7Þ

    D _E 2n    14c

    dknDR 2Lsn   BR ðc 1ÞCos2ð xcnÞ   C

    1 þ  ffiffiffiffiffiPr p  K 1

    0@

    1A

    24

    Sin2ð xcnÞ

     ffiffiffiffiffiPr 

    BR  K

    #  ð8Þ

     Table 1

    Normalized parameters

    Independent parameters Normalizing parameters Normalized parameters

    Length and position   k2p   Normalized length and position

    Plate spacing The sum of plate spacing and thickness Blockage ratio

    Penetration depths Half of the stack plate spacing Normalized penetration depths

    Pressure amplitude Mean pressure Drive ratio

    Temperature difference along the stack Mean temperature of the stack Normalized temperature difference

    Temperature gradient along the stack Critical temperature gradient Normalized temperature gradient

    Power The product of mean pr es sure, s ound velocit y and gas cross -sectional area in the s tack Normalized power

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3589

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    6/14

    In this study, by using the same normalizing parameters as for Eqs.

    (7) and (8), the normalized acoustic power dissipated in a half 

    wavelength resonator is estimated as,

    D _E 2n;r    p8c

     ffiffiffiffiffiPr 

    p   DR 2

      dk

    r hr

      ðc 1Þp

    8c  DR 2

      dk

    r hr

      ð9Þ

    The energy dissipated in the resonator is proportional to the wall

    surface area of the resonator. So to determine the normalizedacoustic power dissipated in a quarter wavelength resonator, the

    above equation should be divided by two.

     3.2. Energy balance

    One purpose of the analyses presented in this section is to cor-

    relate some important thermoacoustic parameters. The required

    heat input to the device is correlated with the total energy flux

    through the engine stack and acoustic energy flux, by applying

    the energy balance on the hot heat exchanger. The desired cooling

    power of the device is correlated with the total energy flux through

    the refrigerator stack and the acoustic energy flux by applying en-

    ergy balance on the cold heat exchanger. The analyses cover all

    possible characteristics that could be assumed for thermoacoustic

    engines and refrigerators. These relationships would also provide a

    better estimation of the engine and refrigerator performance.

    Consider a thermoacoustic engine with its stack located near

    the left pressure antinode of the resonator as illustrated in Fig. 3.

    The energy balance is applied on the control volume outlined with

    the dashed line which encloses the hot heat exchanger. Thus,

    Q h ¼   _H 2;s þ   _H 2;hd   ð10Þwhere   _H 2;hd  is the acoustic power leaving the control volume into

    the hot duct, which is dissipated in the hot duct. If it is assumed that

    the heat generated by the dissipation of the acoustic power in the

    hot duct   ðD _E 2;hdÞ   is rejected to the environment through the hot

    duct walls then,  ð _H 2;hd  ¼ D _E 2;hdÞ  and we have  Q h  ¼   _H 2;s þD _E 2;hd. As

    the hot duct is always a small portion of the resonator, the magni-

    tude of  D _E 2;hd  is very small and can be neglected, thus,

    Q h ¼   _H 2;s   ð11Þ

    This implies that the total energy flux into the engine stack is

    approximately equal to the heat input by the hot heat exchanger.

    It should be noted that Eq. (11) will also be applicable when the en-

    gine stack is placed near the right-end pressure antinode or when

    the resonator walls are insulated.

    The ratio of the acoustic work produced by the engine stack to

    the energy flux delivered to the system by HXh   is defined as the

    thermal efficiency of the engine stack, gth,s

    , expressed as,

    gth;s ¼D _E 2;s;eng

    Q h¼ D

    _E 2n;s;engQ hn

    ð12Þ

    where Q h(Q hn) is defined in Eq.  (11).

    Fig. 4a and b shows two possible configurations of a thermoa-

    coustic refrigerator, i.e. the refrigerator stack located near the left

    pressure antinode or right pressure antinode of the resonator,

    respectively. Note that the acoustic power to the refrigerators

    (either by an engine or a loud speaker) is provided from the left

    side of the resonator in both cases. In other word, the engine stack

    (or loud speaker) is located on the left side of the refrigerator stack.

    A thermoacoustic refrigerator with its stack located near the left

    pressure antinode is illustrated in Fig. 4a. The energy balance is ap-

    plied on the control volume outlined with the dashed line in Fig. 4a

    which encloses the cold heat exchanger. Thus,

    Q c ¼   _H 2;s þ   _H 2;cd   ð13Þwhere Q c  is the desired cooling power,   _H 2;s  is the total energy flow

    towards the stack which is equal to the sum of heat extracted from

    the cold heat exchanger and the heat produced by the acoustic

    power dissipation in the cold heat exchanger, minus the acoustic

    power enter the control volume.   _H 2;cd  is the acoustic power leaving

    the control volume into the cold duct. This acoustic power is dissi-

    pated in the cold duct. If it is assumed that the heat generated by

    the dissipation of the acoustic power in the cold duct ðD _E 2;cdÞ is re-

     jected to the environment through the cold duct walls then,

    _H 2;cd

     ¼D _E 2;cd

      ð14

    Þand,Q c D _E 2;cd ¼   _H 2;s   ð15ÞSince the actual net cooling power is the amount of energy flux re-

    moved from the cold heat exchanger and pumped uphill by the

    stack, the actual cooling power of the device is   Q c D_E 2;cd   in this

    case.

    If the cold duct is insulated then it could be assumed that D _E 2;cdis not rejected to the environment and it appears as a load on the

    cold heat exchanger. This heat leaves the cold duct and enters the

    control volume and flows into the stack.

    Thus,   _H 2;cd  ¼ 0 and,

    Q c

     ¼  _H 2;s

      ð16

    ÞIn other configuration of the thermoacoustic refrigerator, the stackcan be located near the right pressure antinode of the device as

    Fig. 3.  Schematic of a thermoacoustic engine. The control volume is outlined with

    dashed lines which encloses the hot heat exchanger (HXh).

    Fig. 4.   Schematic of a thermoacoustic refrigerator with two possible configurations: (a) Refrigerator stack located near the left pressure antinode, (b) refrigerator stacklocated near the right pressure anitnode. The control volume is outlined with dashed lines which encloses the cold heat exchanger (HX c).

    3590   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    7/14

    shown in  Fig. 4b. Consider the control volume outlined with the

    dashed line which encloses the cold heat exchanger. If it is assumed

    that the heat generated by the dissipation of the acoustic power in

    the cold portion of the resonator is rejected to the environment,

    then from the energy balance, we have,

    Q c ¼   _H 2;s   _H 2;sum   ð17Þ

    where  _

    H 2;sum   is the acoustic power entering the control volumewhich is the sum of the acoustic power to be dissipated by the cold

    heat exchanger, stack, ambient heat exchanger and ambient duct.

    Neglecting the acoustic power dissipated in the ambient duct, and

    defining   D _E 2;t   as the acoustic power dissipated/consumed in the

    cold heat exchanger, stack, ambient heat exchanger, we have,

    Q c ¼   _H 2;s D _E 2;t   ð18ÞThis equation implies that the total energy flux of the stack is the

    cooling power of the system and the acoustic power consumed by

    the stack and the heat exchangers. If the cold duct is insulated,

    D _E 2;cd   is not rejected to the environment through the cold duct

    walls, and it appears as a load on the cold heat exchanger. Thus,

    _H 2;sum

     ¼D _E 2;t

    þD _E 2;cd

      ð19

    Þand,

    Q c ¼   _H 2;s ðD _E 2;t þ D _E 2;cdÞ ð20ÞThe ratio of the cooling power of a thermoacoustic refrigerator to

    the consumed acoustic power by the stack is defined as the coeffi-

    cient of performance of the refrigerator stack.

    COPs ¼ Cooling powerD _E 2;s;ref 

    ¼ Normalized cooling powerD _E 2n;s;ref 

    ð21Þ

    where cooling power is defined in Eqs.   (15), (16), (18) and (20).

    Considering a thermoacoustically-driven thermoacoustic refrigera-

    tor, the acoustic power produced by the thermoacoustic engine

    must be consumed by the thermoacoustic refrigerator and the res-

    onator. That is,

    D _E 2;s;eng ðD _E 2;HXh þ D _E 2;HXa;eng Þ¼ D _E 2;s;ref  þ D _E 2;HXc þ D _E 2;HXa;ref  þ D _E 2;r   ð22Þ

    The left hand side is the net acoustic power output of the thermoa-

    coustic engine. That is, the total acoustic power available for the

    refrigeration purpose. The first three terms on the right hand side

    are the total acoustic power consumed by the refrigerator stack

    and its exchangers and the last term is the dissipated acoustic

    power in the resonator. The above equation could be summarized

    as,

    D _E 2;t;eng ¼ D _E 2;t;ref  þ D _E 2;r   ð23Þ

    Thus,D _E 2;pro ¼ D _E 2;con   ð24Þ

    where   D _E 2;pro   is the acoustic power produced in the engine and

    D _E 2;con   is the acoustic power consumed in the refrigerator and res-

    onator. In the normalized form, the above equation can be ex-

    pressed as,

    D _E 2n;pro ð AP maÞeng ¼ D _E 2n;con ð AP maÞref    ð25ÞParameters A  and  P m  are the same for both engine and refrigerator

    whereas, the speed of sound is different in both stacks due to the

    difference in the mean temperatures of the stacks. Eq. (25) can fur-

    ther be simplified as,

    D _E 2n;pro ¼   aref aeng

    D _E 2n;con   ð26Þ

     3.3. Entropy balance

    To determine the entropy generation within a thermoacoustic

    device, the entropy balance and energy balance are applied on

    two control volumes. The first control volume (system I) is the

    acoustic power producing system which consists of the hot heat

    exchanger, engine stack and the engine ambient heat exchanger,

    as outlined in Fig. 5a with the dashed line. The second control vol-

    ume (system II) is the acoustic power consuming system whichconsists of the resonator, cold heat exchanger, refrigerator stack

    and ambient heat exchanger, as outlined in Fig. 5b by the dashed

    line. Since the entropy change of a steady state control volume is

    zero, the second law of thermodynamics indicates that the entropy

    leaving the control volume must equal the sum of the entropy

    entering the control volume and the entropy generation within

    the control volume. It is useful to mention that there is no entropy

    associated with energy transfer as work [20].

    Following equations show the energy balance on systems I and

    II, respectively.

    Q h ¼Q a;eng þ D _E 2;pro   ð27ÞQ c

    þD _E 2;con

     ¼Q a;ref 

     þQ r

      ð28

    Þwhere Q r represents the amount of dissipated acoustic power in theresonator leaving through the resonator’s wall in the form of heat

    energy at the ambient temperature.

    Following equations show the entropy balance on systems I and

    II, respectively.

    Q a;engT a

    ¼Q hT h

    þ  _S gen;eng   ð29ÞQ a;ref T a

    þQ rT a

    ¼ Q cT c

    þ  _S gen;r;ref    ð30Þ

    By substituting the values of  Q a,eng and  Q a,ref  from Eqs. (27) and (28)

    into Eqs. (29) and (30), respectively, we get,

    T a

    _S gen;eng ¼

    Q h   1

    T a

    T h D

    _E 2;pro  ð

    31

    ÞT a _S gen;r;ref  ¼ Q c   1

    T aT c

    þ D _E 2;con   ð32Þ

    Fig. 5.   Two thermodynamic systems outlined by dashed lines, (a) system I, acoustic power producing system (engine), (b) system II, acoustic power consuming system(resonator and refrigerator).

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3591

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    8/14

    Considering a thermoacoustically-driven thermoacoustic refrigera-

    tor (TADTAR), the total entropy generation within the TADTAR 

    could be estimated by adding the entropy generation in the two

    systems (since the two systems form a TADTAR). Thus,

    T a  _S gen;t ¼ T a  _S gen;eng þ T a  _S gen;r;ref ¼   Q h   1

    T aT h þ Q c   1

    T aT c

    þ ½D _E 2;con D _E 2;pro ð33ÞTo satisfy Eq. (24) which implies that the produced acoustic power

    by the engine stack must be equal to the consumed acoustic power

    by the other components of the device, the engine characteristics

    must be selected to make the last term on the right hand side of 

    Eq.  (33) equal to zero. Thus, the first term on the right hand side

    of Eq. (33) shows the total entropy generation within the TADTAR.

    To have the maximum efficiency, the engine characteristics must be

    selected to make the entropy generation minimum. For an acousti-

    cally-driven thermoacoustic refrigerator (ADTAR), the total entropy

    generation could be determined by applying Eq.  (32).

    In the following subsections, the developed design and optimi-

    zation algorithm is explained in detail and, an example is pre-

    sented as a case study to demonstrate the functioning of thealgorithm. Some important issues related to the designing are also

    discussed.

     3.4. Design and optimization algorithm

    The algorithm proposed by Wetzel and Herman [13] to design

    acoustically-driven thermoacoustic refrigerators was used as the

    base model to develop a new systematic comprehensive algorithm

    in this study. The developed algorithm can be used to design and

    optimize not only thermoacoustically-driven thermoacoustic

    refrigerator but also individual thermoacoustic engines and

    acoustically-driven thermoacoustic refrigerators. Furthermore,

    the present algorithm includes new features for designing an

    acoustically-driven thermoacoustic refrigerator by incorporatingcorrelations between different design parameters based on the

    energy balance, that were not available in the previous studies.

    The complete design algorithm is shown in Fig. 6. As mentioned

    earlier, meeting the required cooling power at the desired cooling

    temperature and at the given hot heat temperature while rejecting

    someheat to the environment can bedefined as the goals of design-

    ing a thermoacoustically-driven thermoacoustic refrigerator.

    The design procedure starts with the refrigerator section of the

    device followed by the resonator and the heat engine. As a first

    step, the designer must pick the working gas, blockage ratio, ther-

    mal penetration depth and drive ratio of the device (similar to

    previous studies). Using the values of the given cooling tempera-

    ture (i.e. the temperature of HXc), heat input temperature (i.e.

    the temperature of HXh) and, the surrounding ambient tempera-ture (i.e. the temperature of HXa), the mean temperatures of the

    refrigerator and engine stacks can be calculated. The thermophys-

    ical properties of the working gas are then computed in the refrig-

    erator section of the device (and the resonator) and engine section

    of the device based on the refrigerator and engine stack mean tem-

    peratures, respectively.

    As mentioned above, the designing process starts with the

    refrigerator. In the energy balance section, two configurations are

    presented for a thermoacoustic refrigerator (Fig. 4a and b). For each

    configuration, two conditions are presented, i.e. cold duct insulated

    and uninsulated. The relationship between cooling power, total en-

    ergy flux to the stack and the acoustic power for all cases are also

    presented (see Eqs. (15), (16), (18) and (20)). The designer must

    select the desired configuration and apply the appropriate energy

    balance on HXc.

    By plotting COPs  and COPR s  as functions of  xcn,ref ,  Lsn,ref ,  DT n,ref ,

    the normalized refrigerator stack length and position at the desired

    normalized temperature difference is selected. The resonator

    cross-sectional area (and then the resonator hydraulic radius)

    can be determined by,

     Ar ¼   Q cQ cn

    P m

    a

    BR 

      ð34Þ

    In the next step, the normalized acoustic power dissipated in the

    refrigerator stack  ðD _E 2n;s;ref Þ is determined by applying Eq. (8). This

    equation is also used to determine the normalized acoustic power

    dissipated in heat exchangers ðD _E 2n;HXc  þ D_E 2n;HXa;ref Þ by substituting

    the normalized heat exchanger length and position in Eq.  (8) and

    assuming no temperature gradient along the exchangers (i.e. along

    the direction of acoustic wave propagation). The value   D _E 2n;t;ref which is the sum of the above mentioned values can then be

    determined.

    To design the resonator, select the resonance frequency of the

    thermoacoustic device. Compute the thermophysical properties

    of the working gas in the resonator based on stack mean tempera-

    ture of the refrigerator. Calculate the thermal penetration depth at

    the resonator’s wall. The length of the resonator can be set equal tohalf wavelength or quarter wavelength of the resonant standing

    wave in the device. Determine the normalized acoustic power dis-

    sipated in the resonator ðD _E 2n;rÞ by using Eq. (9).

    The resonance frequency of the acoustic standing wave is an

    important design parameter. Although it has been selected at this

    stage, it could be modified afterwards if necessary. Higher reso-

    nance frequency results in lower penetration depth i.e. small plate

    spacing in the stack which increases the manufacturing challenge,

    however, higher resonance frequency increases the power density

    and reduces the length of the resonator. In the next step, the total

    consumed acoustic power ðD _E 2n;conÞ can be computed by summing

    the dissipated acoustic power in the refrigerator stack, its heat

    exchangers  ðD _E 2n;t;ref Þ and the resonator  ðD _E 2n;rÞ.

    The purpose of the heat engine is to produce the acoustic powerthat is consumed by the device. Once the total consumed acoustic

    power is computed ðD _E 2n;conÞ, the acoustic power to be produced by

    the engine   ðD _E 2n;proÞ   can be estimated by using Eq.  (26). This

    parameter serves as a basis to design the heat engine which can

    meet the given requirements.

    To design the engine stack of the device, estimate the stack effi-

    ciency based on the energy balance applied on the hot heat ex-

    changer using Eq.   (12). The results from this energy balance

    would also be used in estimating the heat input to the engine. In

    the next step, the length and position of the engine stack must

    be selected that could produce the estimated acoustic power at

    the desired heat exchangers’ temperature while having the maxi-

    mum possible efficiency. The appropriate length and position for

    the engine stack can be selected by plotting the engine thermalefficiency (gth), the engine normalized thermal efficiency (gthn)and   D _E 2n;pro   as functions of   xcn,eng,   Lsn,eng,   D T n,eng. At this stage,

    the entropy balance described in the previous section is applied

    on the device to evaluate the overall entropy generation within

    the thermoacoustic device (Eq.  (33)). This analysis is useful to

    examine the variation of the generated entropy as a function of 

     xcn,eng,  Lsn,eng, so the engine stack position and length are selected

    to minimize the entropy generation while providing the required

    acoustic power at the given temperatures of the heat exchangers.

    The point of the minimum entropy generation is the same as the

    point of the maximum efficiency of the device.

    After selecting the optimized engine stack length and position,

    estimate the amount of heat input to the hot heat exchanger to

    produce the estimated acoustic power at the desired temperatures.

    The heat input can be computed by,

    3592   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

    http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-http://-/?-

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    9/14

    Q h ¼ Q hn ð Ar BR  P m aengÞ ð35ÞIn the last step, the plate thickness and plate spacing of engine and

    refrigerator stacks are estimated.

    As mentioned earlier, the present algorithm can also be used to

    design an acoustically-driven thermoacoustic refrigerator. In this

    case, the same procedure and steps are used to calculate the total

    acoustic power consumed by the refrigerator stack, heat exchang-

    ers and the resonator   ðD _E 2n;conÞ. A loud speaker is then selected

    based on this total acoustic power to run the apparatus.

    It is useful to mention that this study is more comprehensive

    compared to previous studies for designing and optimizing ther-moacoustic refrigerators;since, it usesthe cooling power andenergy

    flux relations for different configurations of the thermoacoustic

    refrigerator.

     3.5. Case study

    A case study is presented to demonstrate the working of the

    developed procedure. The case study comprised of designing a

    TADTAR with 30 W of cooling power at the desired cooling temper-

    ature of 277 K and the desired hot heat exchanger temperature of 

    623 K. The ambient heat exchangers are assumed to operate at

    300 K. Helium at a mean pressure of 700 kPa is selected as theworking gas, which is one of the recommended gases for thermoa-

    Fig. 6.  Schematic of developed design and optimization algorithm for thermoacoustically-driven thermoacoustic refrigerators.

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3593

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    10/14

    coustic devices [21]. The other parameters required at the begin-

    ning of the designing procedure are set as follows; BR = 0.8,

    dkn = 0.66 and DR = 0.03, which are consistent with the previous

    studies   [17,22]. Based on the given temperatures for hot, cold

    and ambient heat exchangers, the mean temperature of the refrig-

    erator stack is 288.5 K (DT n,ref  = 0.08) and the mean temperature of 

    the engine stack is 461.5 K (DT n,eng = 0.7).

    As mentioned in the previous section, the design procedurestarts with the designing of the refrigerator section. The second

    configuration of the refrigerator is selected with no insulation on

    the cold duct. That is, the refrigerator stack is located near the right

    pressure antinode of the resonator (see Fig. 4b), and the dissipated

    acoustic power in the resonator is rejected to the environment

    through the resonator’s wall.

    The next step is the selection of the normalized refrigerator

    stack length and position by plotting COPs and COPR s as functions

    of   xcn,ref ,   Lsn,ref   at the desired normalized temperature difference.

    Fig. 7a and b show the variation of COPs and COPR s as the function

    of  Lsn,ref  at different xcn,ref  and DT n,ref  = 0.08. The plots show that by

    shifting the stack center away from the pressure antinode (i.e.

    increasing x cn,ref ), the stack length must be increased to have the

    performance peak. However, the peak magnitude decreases with

    increasing xcn,ref . By selecting the length of the stack to have the

    performance peak, two problems arise. First, as the figures show,

    the COPs values are very sensitive to the stack length near the peak.

    A slightly smaller stack length causes a sharp decrease in the per-

    formance of the refrigerator. Second, the apparatus does not per-

    form as a refrigerator at higher values of the normalized

    temperature differences (discussed in a later section). Taking into

    considerations these issues, it is decided to assume   xcn,ref  = 0.11

    and   Lsn,ref  = 0.035 (COPs = 4.47, COPR s  = 0.37). The cross-sectional

    area of the resonator is computed by using Eq.   (34), which for

    the present case is equal to 0.0123 m2.

    The variation of the normalized consumed acoustic power

    ðD _E 2n;conÞ is plotted versus the resonance frequency at the selected

    specifications of the refrigerator stack in   Fig. 8. The values of 

    D _E 2n;con  are computed using Eqs.   (8) and (9), as described in theprevious section. The plot shows that at a given refrigerator stack

    temperature difference, the consumed acoustic power decreases

    by increasing the resonance frequency. As the resonance frequency

    increases, the viscous and thermal penetration depths decreases

    causing the acoustic power dissipated in the resonator to decrease.

    Thus, it is desirable to have a thermoacoustic device operating at

    higher resonance frequency. However, higher frequency results

    in lower peak-to-peak displacement of the gas particles which

    could affect the performance of the heat exchangers [2].

    In the present study, the resonance frequency of 400 Hz is se-

    lected. For this frequency, the length of the resonator based on

    the half wavelength of the acoustic standing wave is estimated

    to be 1.25 m. The normalized consumed acoustic power and nor-

    malized acoustic power to be produced by the engine are esti-

    mated to be   D _E 2n;con  ¼ 3:2 106 and   D _E 2n;pro  ¼ 2:5510

    6,

    respectively.

    The engine stack is considered to be located near the left pres-

    sure antinode. In the next step, the length and position of the en-

    gine stack is selected for the given normalized temperature

    difference. The selection is done from the graphs of   gth,s,   gthn,sand  D _E 2n;pro  versus the normalized engine stack length (Lsn,eng) at

    different values of   xcn,eng   for the given  DT n,eng, that are obtained

    by applying the energy balance on the hot heat exchanger.  Fig. 9

    shows the variations of   gth,s,   gthn,s   and   D _E 2n;pro   versus   Lsn,eng   at

    Fig. 7.  Normalized length of the refrigerator stack (Lsn,ref ) versus (a) coefficient of performance of refrigerator stack (COPs), (b) coefficient of performance relative to Carnotcycle (COPR s), at different normalized stack center positions ( xcn,ref ) at  DT n,ref  = 0.08.

    Fig. 8.   Normalized consumed acoustic power   ðD _E 2n;conÞ   versus the resonance

    frequency ( f ) at  xcn,ref  = 0.11, Lsn,ref  = 0.035 and DT n,ref  = 0.08.

    3594   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

    http://-/?-http://-/?-

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    11/14

    different xcn,eng  and at  DT n,eng = 0.7.  Fig. 9a shows that at a given

    stack position, the acoustic power produced by the engine de-

    creases with an increase in the stack length, whereas, as a given

    stack length, the acoustic power produced by the engine increases

    as the stack moves away from the pressure antinode. For the esti-

    mated value   D_

    E 2n;pro  ¼ 2:55106

    in the present study, severalcombinations of the stack length and position are available from

    short stack close to the pressure antinode to long stack away from

    the pressure antinode. Based on the combination of   Lsn,eng   and

     xcn,eng   that can produce the required acoustic power, the corre-

    sponding values of  gth,s   and  gthn,s  can be estimated from  Fig. 9band c, respectively, to evaluate the stack performance. Fig. 9 also

    shows that for a specified stack center position there is not more

    than one stack length that could produce the required acoustic

    power at the desired normalized temperature difference. The re-sults show that at the combination xcn,eng = 0.14 and  Lsn,eng = 0.09,

    the stack performance is the best i.e., gth,s = 18% and  gthn,s = 0.347.Fig. 9a shows that certain combinations of  Lsn,eng and  xcn,eng cannot

    produce the required acoustic power. It also shows that below a

    certain value of  xcn,eng, there is no length of stack that could pro-

    duce the required acoustic power, which in the present case is

     xcn,eng 6 0.06. Thus, for a specified thermoacoustic refrigerator,

    there could be an engine stack with specified position and length

    that produces the required acoustic power with the maximum pos-

    sible efficiency. Thus, the device must generate the least entropy at

    these specifications.

    To check if the total entropy generation  ð _S gen;tÞ   in TADTAR is

    minimum,   _S gen;t   is computed using Eq. (33) at the selected combi-

    nations of   xcn,eng   and   Lsn,eng  at which the required   D _E 2n;pro   is ob-

    tained. The   _S gen;t   is plotted as a function of   xcn,eng   in   Fig. 10. For

     xcn,eng 6 0.06, there is no engine stack length that could produce

    the required acoustic power of the system. Fig. 10 shows that the

    device would generate minimum entropy by placing the engine

    stack at xcn,eng = 0.14 with the corresponding value of  Lsn,eng = 0.09,

    which confirms the above mentioned discussion. Thus, it can be

    concluded that while selecting the position and length for the en-

    gine stack that produces the required acoustic power, the designer

    should confirm that the entropy generation of the device is mini-

    mum at the selected specifications.

    In the final step, the amount of heat input to the hot heat ex-

    changer (Q h) to produce the required acoustic power at the desired

    temperatures is estimated using Eq. (35). For the present TADTAR,

    Q h = 162.2 W. Thus, for the thermoacoustically-driven thermoa-

    coustic refrigerator to produce the cooling power of 30 W,162.2 W of heat input is required.

    The plate thickness and plate spacing of the refrigerator stack

    are estimated to be approximately 0.10 mm and 0.42 mm, respec-

    tively. The plate thickness and plate spacing of the engine stack are

    estimated to be about 0.16 and 0.63 mm, respectively.

    Fig. 10.   Total entropy generation in the device   ð_S gen;tÞ   versus normalized stack

    center positions ( xcn,eng).

    Fig. 9.  Normalized length of the engine stack (Lsn,eng) versus (a) normalized acoustic

    power produced   ðD _E 2n;proÞ, (b) thermal efficiency of engine stack (gth,s), (c)normalized thermal efficiency of engine stack (gthn,s), at different normalized stackcenter positions ( xcn,eng) at  DT n,eng = 0.7.

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3595

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    12/14

     3.6. Effects of stack temperature difference (DT n) on the design of a

    thermoacoustic device

    As described in the design algorithm, the designer must select

    the temperatures of all heat exchangers (i.e. stack temperature dif-

    ference) at the beginning of the design procedure. The equations

    show that the stack temperature difference (DT n) is an important

    design parameter and has a significant influence on the perfor-mance of the respective engine or refrigerator. Therefore, it is

    important for a designer to have a good understanding of the influ-

    ence of DT n on the performance of the device. Since the design pro-

    cedure is based on selected values of   DT n   for engine and

    refrigerator, the effect of DT n on their performance cannot be eval-

    uated in the previous sections. In this section, the effects of DT n on

    the overall design of the device are discussed in detail.

    The impact of the refrigerator stack temperature difference

    (DT n,ref ) on the performance of the refrigerator stack is illustrated

    in Fig. 11a and b, where the variations of COPs and COPR s are plot-

    ted as a function of  Lsn,ref  at different DT n,ref  and at a given value of 

     xcn,ref . The figure shows that the peak performance of the refriger-

    ator stack reduces by increasing the stack temperature difference.

    The trends in the given figure also indicate that for a given stack

    temperature difference, the COP drops to zero if the length of the

    stack is lower than a certain value. That is, for a stack to operate

    as a refrigerator, the length of the stack should be higher than a

    cutoff value (also see Eq. (6)). The figure shows that the cutoff va-

    lue of the stack length increases with decreasing the stack temper-

    ature difference. In the case study, at   DT n,ref  = 0.08, the best

    performance of the refrigerator stack is at Lsn,ref  = 0.025. However,

    if the temperature difference is increased, the stack may not per-

    form as a refrigerator. This could happen when developing the ac-

    tual device as the actual stack temperature difference may vary

    from its designed value. Therefore, it is safer to select the stack

    length slightly larger than that correspond to the peak perfor-

    mance. Therefore, in the case study the length of the refrigerator

    stack was selected as Lsn,ref  = 0.035.

    The influence of stack temperature difference on the resonancefrequency is shown in Fig. 12, where the variation of the normal-

    ized consumed acoustic power ðD _E 2n;conÞ is plotted versus the nor-

    malized refrigerator stack temperature difference (DT n,ref ) at

    different resonance frequencies at   xcn,ref  = 0.11 and   Lsn,ref  = 0.035.

    At a given resonance frequency, the consumed acoustic power de-

    creases with an increase in the stack temperature difference. As the

    temperature difference along the refrigerator stack increases, the

    thermal penetration depth decreases, causing a reduction in the

    acoustic power consumed in the stack.

    The influence of the engine stack temperature difference on the

    performance of the engine stack is shown in  Fig. 13. In this figure,

    the variations of  gth,s, gthn,s and  D _E 2n;pro  are plotted versus the nor-malized engine stack length (Lsn,eng) at different values of D  T n,eng.

    This figure shows that the peak efficiency of the engine stack in-

    creases by increasing the normalized temperature difference along

    the engine stack. The trends in  Fig. 13 also indicate that for a given

    stack temperature difference, the stack efficiency drops to zero if 

    the length of the stack is greater than a certain value. That is, for

    a stack to operate as an engine, the length of the stack should be

    lower than a cutoff value (also see Eq.  (6)). The figure shows that

    the cutoff value of the stack length decreases with decreasing the

    stack temperature difference.

    The influence of stack temperature difference for engine and

    refrigerator on the heat input to the device is shown in  Fig. 14.

    The heat input is plotted as a function of normalized engine stack

    Fig. 11.  Normalized length of the refrigerator stack (Lsn,ref ) versus (a) coefficient of performance of refrigerator stack (COPs), (b) coefficient of performance relative to Carnotcycle (COPR s), at different values of normalized refrigerator stack temperature difference (DT n,ref ) at  xcn,ref  = 0.11.

    Fig. 12.  Normalized consumed acoustic power  ðD _E 2n;conÞ versus normalized refrig-

    erator stack temperature difference (DT n,ref ) at different resonance frequencies, at

     xcn,ref  = 0.11, Lsn,ref  = 0.035.

    3596   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    13/14

    temperature difference at different normalized refrigerator stack

    temperature difference at the specified engine and refrigerator

    stack positions and lengths. The figure shows that by increasing

    D

    T n,eng   or  D

    T n,ref   or both, the required heat input to the deviceincreases.

    4. DeltaE

    The computer code DeltaE can be used to simulate the devices

    designed and optimized by the procedure presented in this study.

    DeltaE solves the one-dimensional wave equation in gas or liquid,

    based on the low amplitude acoustic approximation in user de-

    fined geometries   [23]. The desired parameters initially selected

    and the parameters computed by using the design and optimiza-

    tion algorithm developed in this study are summarized in  Table

    2. Also presented in the table are the values obtained from DeltaE

    for comparison. A good agreement is observed between the devel-

    oped procedure and the computer code DeltaE. Although the

    parameters calculated by the computer code DeltaE are slightlydifferent from those of estimated by the developed procedure, it

    is reasonable to say that the developed procedure can serve as a

    great tool to design and optimize thermoacoustic devices since

    designing a TADTAR by using the computer code DeltaE to meet

    the designer’s requirements requires tremendous numbers of trials

    and errors making this job tedious. The small differences between

    the two approaches are mainly due to the assumptions that were

    made to linearize and simplify the governing equations to develop

    the design and optimization procedure. The inaccurate expression

    used to estimate the temperature difference between the metal

    and working gas in the heat exchangers in the computer code

    DeltaE could be another reason for the deviations  [23]. One or

    Fig. 13.  Normalized length of the engine stack (Lsn,eng) versus (a) thermal efficiency

    of engine stack (gth,s), (b) normalized thermal efficiency of engine stack (gthn,s), (c)normalized acoustic power produced ðD _E 2n;proÞ, at different normalized engine stack

    temperature difference (DT n,eng) at xcn,eng = 0.09.

    Fig. 14.  Heat input (Q h) to the device versus normalized engine stack temperature

    difference (DT n,eng) for different values of normalized refrigerator stack temperature

    difference (DT n,ref ).

     Table 2

    Comparison between results from present algorithm and DeltaE simulations

    Present algorithm DetlaE

     f    400 402.7

    T HXh   623 630

    T HXa;eng   3 00 300.5

    DT n,eng   0.7 0.708

    T HXc   277 277

    T HXa;ref    300 303.3

    DT n,ref    0.08 0.09

    Q h   164.7 166.1

    Q c   30 30

    D _E 2;s;eng   29.7 32.9

    D _E 2;s;ref    6.72 6.9

    gth,s (%) 18 19.8COPs   4.47 4.3

    Overall efficiency (%) 18.5 18.1

    H. Babaei, K. Siddiqui / Energy Conversion and Management 49 (2008) 3585–3598   3597

  • 8/19/2019 Babaei Design and Optimization of Thermoacoustic Devices

    14/14

    more parameters such as stack center position, stack length or

    resonator cross-sectional area can be adjusted to meet the desired

    values in DeltaE.

    5. Conclusion

    Thermoacoustic devices operate by the energy conversion

    between heat and sound, and have no harmful effects on the envi-

    ronment. The designing of thermoacoustic devices involves signif-

    icant technical challenges. In the present study, a comprehensive

    design and optimization algorithm is developed for designing ther-

    moacoustic devices. The unique feature of the present algorithm is

    its ability to design thermoacoustically-driven thermoacoustic

    refrigerators that can serve as sustainable refrigeration systems.

    In addition, new features based on the energy balance are also in-

    cluded to design individual thermoacoustic engines and acousti-

    cally-driven thermoacoustic refrigerators. The algorithm is based

    on the simplified linear thermoacoustic model. It includes different

    correlations based on the energy balance for different device con-

    figurations. Another important feature of the algorithm is the

    implementation of the entropy balance on the device to refine

    the optimization process. A step-by-step design and optimization

    procedure is described which is followed by a case study inwhich a thermoacoustically-driven thermoacoustic refrigerator is

    designed and optimized to demonstrate the working of the algo-

    rithm. The results from the algorithm are in good agreement with

    that obtained from the computer code DeltaE.

     Acknowledgement

    This research is funded by a grant from the Concordia Univer-

    sity to Kamran Siddiqui.

    References

    [1] Rott N. Thermoacoustics. Adv Appl Mech 1980;20:135–74.[2] Swift GW. Thermoacoustic engines. J Acoust Soc Am 1988;84:1145–79.

    [3] Garret SL, Adeff JA, Hofler TJ. Thermoacoustic refrigerator for space application. J Thermophys Heat Transfer 1993;7:595.

    [4] Hofler TJ. Thermoacoustic refrigerator design and performance. DissertationUniversity of California, San Diego; 1986.

    [5] Tijani MEH, Zeegers JCH, De Waele ATAM. Construction and performance of athermoacoustic refrigerator. Cryogenics 2002;42:59–66.

    [6] Symko OG, Abdel-Rahman E, Kwon YS, Emmi M, Behunin R. Design anddevelopment of high-frequency thermoacoustic engines for thermalmanagement of microelectronics. Microelectron J 2004;35:185–91.

    [7] Hatazawa M, Sugita H, Ogawa T, Seo Y. Performance of a thermoacoustic sound

    wave generator driven with waste heat of automobile gasoline engine. Trans Jpn Soc Mech Eng 2004;70:292–9.

    [8] Adeff JA, Hofler TJ. Design and construction of a solar poweredthermoacoustically driven thermoacoustic refrigerator. J Acoust Soc Am2000;107:L37–42.

    [9] Babaei H, Siddiqui K, Chishty WA. Sustainable thermoacoustic refrigerationsystem for gas turbine power plants. In: 17th symposium of industrialapplication of gas turbines (IGAT); Banff, Canada; 2007.

    [10] http://www.ecn.nl/fileadmin/ecn/units/eei/Onderzoeksclusters/Restwarmtebenutting/b-07-007.pdf .

    [11] Johnson VH. Heat generated cooling opportunities in vehicles. Society of Automotive Engineers; 2002.

    [12] Zoontjens L, Howard C, Zander A, Cazzolate B. Feasibility study of anautomotive thermoacoustic refrigerator. In: Proceedings of acoustics.Busselton, Australia, November 9–11; 2005.

    [13] Wetzel M, Herman C. Design optimization of thermoacoustic refrigerators. Int J Refrig 1997;20:3–21.

    [14] Tijani MEH, Zeegers JCH, De Waele ATAM. Design of thermoacousticrefrigerators. Cryogenics 2002;42:49–57.

    [15] Swift GW. Thermoacoustics: a unifying perspective for some engines andrefrigerators. The Acoustical Society of America, NY: Melville; 2002.

    [16] Olsen JR, Swift GW. Similitude in thermoacoustics. J Acoust Soc Am1994;95:1405–12.

    [17] Worlikar A, Knio O. Numerical simulation of a thermoacoustic refrigerator. JComput Phys 1996;127:424–51.

    [18] Sergeev SI. Fluid oscillations in pipes at moderate Reynolds number. Fluid Dyn1966;1:121–2.

    [19] Merkli P, Thomann H. Transition to turbulence in oscillating pipe flow. J FluidMech 1975;68:567–76.

    [20] Cengel YA, Boles MA. Thermodynamics: an engineering approach. Fourthedition in SI units. McGraw Hill; 2002.

    [21] Belcher JR, Slaton WV, Raspet R, Bass HE, Lightfoot J. Working gases inthermoacoustic engines. J Acoust Soc Am 1999;105:2677–84.

    [22] Tijani MEH, Zeegers JCH, De Waele ATAM. The optimal stack spacing forthermoacoustic refrigeration. J Acoust Soc Am 2002;112:128–33.

    [23] Ward WC, Swift GW. Design environment for low-amplitude thermoacousticengine. J Acoust Soc Am 1994;95:3671–4.

    3598   H. Babaei, K. Siddiqui/ Energy Conversion and Management 49 (2008) 3585–3598

    http://www.ecn.nl/fileadmin/ecn/units/eei/Onderzoeksclusters/Restwarmtebenutting/b-07-007.pdfhttp://www.ecn.nl/fileadmin/ecn/units/eei/Onderzoeksclusters/Restwarmtebenutting/b-07-007.pdfhttp://www.ecn.nl/fileadmin/ecn/units/eei/Onderzoeksclusters/Restwarmtebenutting/b-07-007.pdfhttp://www.ecn.nl/fileadmin/ecn/units/eei/Onderzoeksclusters/Restwarmtebenutting/b-07-007.pdf