You are on page 1of 8

Thermoacoustic refrigeration for ice cream sales

Matthew E. Poese, Robert W. M. Smith, Steven L. Garrett, Rene van Gerwen,* and Pete Gosselin**
*

The Penn State University Applied Research Laboratory, PO Box 30, State College, PA 16804 Unilever Engineering Excellence Team, PO Box 114, 3130 AC Vlaardingen, The Netherlands ** Ben & Jerrys Homemade, Inc., 30 Community Dr., So. Burlington 05403

1. INTRODUCTION
Nearly five years ago, there was a great deal of excitement in the thermoacoustics community after the announcement by Backhaus and Swift [1] of the creation of a regenerator-based thermoacoustic-Stirling heat engine that showed a measured 50% efficiency improvement over the efficiency of earlier standing wave stack-based thermoacoustic engines. (For a brief overview of the differences between stack-based and regenerator-based thermoacoustic engines and refrigerators, see the introduction section of Garretts resource letter [2] or Garrett and Backhauss introductory article [3].) Application of their ideas to electricallydriven thermoacoustic refrigerators in order to realize efficiency gains over previous electrically-driven stack-based thermoacoustic refrigerator technology [4, 5, 6] was intriguing and has proved to be fertile ground for innovation. To generate power using a regenerator, Backhaus and Swift [7] created an acoustic phasing network that could present the high acoustic impedance found in standing wave resonators but with traveling-wave phasing between pressure and volume velocity within the regenerator. In that engine, their acoustical network produced a toroidal gas path that allowed acoustically-induced streaming [8] that had to be suppressed by the introduction of a jet pump to produce a static pressure gradient sufficient to oppose the streaming. That engine also used a standing-wave resonator that was about 1/3 of a wavelength long to provide the required high acoustic impedance at the acoustical network location. The prototype described in this article is an adaptation of the ideas described by [7] although several departures are made from the way in which they implemented the critical elements of the acoustic phasing network and resonator that presents a large acoustic impedance to the regenerator. In the prototype described here, the acoustic phasing network is called the vibro-mechanical multiplier since it is not purely acoustic, and the resonator topology is also a hybrid acousto-mechanical implementation called bellows bounce [9]. This first regenerator-based, low-lift thermoacoustic-Stirling refrigerator prototype should be able to be combined with an ice cream sales cabinet, to replace a vapor compression system by a safe and environmentally benign alternative. The prototype has sufficient cooling capacity to maintain ice cream at the proper serving temperature of 18 C in a 200-liter sales/storage cabinet, and has a size appropriate to be integrated into the cabinet. The US based ice cream company Ben & Jerrys Homemade, Inc. and its mother-company the AngloDutch multinational Unilever, being major end users of ice cream cabinets, funded and cooperated in this phase of research, and provide a bridge between the academic research on thermo-acoustic refrigeration and the commercial development.

2. PROTOTYPE DESCRIPTION
Figure 1 shows the parts of the prototype refrigerator in a section view on the left and a solidmodel cutaway on the right. During operation, the machine is inverted from the orientation

shown in Figure 1 to suppress buoyancy-driven thermal convection. As the objective of the prototype is to cool a 200 l ice cream sales cabinet, the prototype can deliver 120 W of effective cooling capacity at a cold load temperature of 24.6 C and an exhaust temperature of 34 C. These design conditions make it possible to combine the thermoacoustic prototype with secondary distribution systems (e.g. pumped glycol loop), and keep the ice cream in the cabinet at a maximum product temperature of 18 C in an ambient air condition of 25 C.

Figure 1: Prototype Schematic. A cut-away solid model of the entire machine showing a cold heat exchanger (dark grey) that is identical to the hot heat exchanger and contained within a thermallyinsulating Ultem plastic plate (brown). The cold heat exchanger plate is in contact with the platform plate (yellow) that contains the regenerator, sensor signal lines and provides attachment surface for the exhaust (ambient) heat exchanger. A second thermally-insulating Ultem plastic plate (green, above platform plate) provides the contoured plenum space that directs the oscillating cold helium gas in and out of the thermal buffer spaces (windows) through the platform. A solid stainless steel plate (dark grey, top of cutaway drawing) is used to seal the platform, cold heat exchanger plate, and plenum plate to the pressure vessel (green) and provide the force necessary to resist the 10 atmospheres of internal helium gas pressure. Enclosing the hot heat exchanger is the vibromechanical multiplier comprised of the compliance volume within the multipliers cylinder (orange) that is terminated by an ordinary loudspeaker cone (purple, labeled as multiplier cone). Directly below the multiplier cone is the power piston cone (light green) that is attached to the bellows (gray). The moving-magnet linear motor, which moves the power piston cone, is shown as a gray rectangle with yellow straps. It is attached to the bottom plate (black) that forms the lower boundary of the pressure vessel. The cylindrical portion of the pressure vessel is shown in green.

2.1 Bellows Bounce Resonator


The moving-magnet loudspeaker is operated at the acousto-mechanical resonance frequency, fd, set by the stiffness of the gas contained inside the bellows and the moving mass of the loudspeaker, bellows and bellows piston. The optimum operating frequency to maximize the fatigue life of the bellows is about 100 Hz. This frequency corresponds to the 1/4 standing wave resonance frequency for compressional waves in the metal bellows [10]. Once the

machine has been constructed, the acousto-mechanical resonance frequency can be tuned to near 100 Hz by adjusting either the mean pressure of the helium gas or by the addition/subtraction of mass on the piston. The wavelength in room temperature helium gas at this frequency is about 10 m, and the effective bellows length is about 17 cm, which makes this lumped-element acousto-mechanical resonator no more than 1/50 of a wavelength. Because of this compactness, the pressure within the bellows volume (in the absence of the feedback network) is spatially uniform. We call this a bellows bounce resonator because the moving mass (comprised of the piston, the moving mass of the motor and the moving mass of the bellows) bounces against stiffness provided by the pressurized helium contained within the bellows. The use of a solid material mass instead of hydrodynamic mass (like that in a purely acoustic resonator) eliminates the nonlinear dissipative effects of turbulence and entry/exit losses (also called minor losses in steady-flow piping networks) present in most pure acoustic resonators driven at high amplitudes [6].

2.2 Vibromechanical Multiplier


The feedback network is comprised of an inertial element and a compliant element that together create a Helmholtz resonator [11]. In this incarnation, the feedback network is called a vibromechanical multiplier; the compliance is provided by the volume of gas enclosed within the multiplier and the inertial element is provided by a conventional 5 1/4 loudspeaker cone intended for audio playback application. This cone is attached to one end of the cylindrical multiplier wall by the outside rim of the Santoprene surround that comes with the cone.1 The resonance frequency of the vibromechanical multiplier containing 20 grams of helium at 10 atmospheres (absolute) of static pressure is approximately 330 Hz, a factor of three greater than the driving frequency. The pressure enhancement inside of the vibromechanical multiplier compared to outside (pH / p1) is 8.5%. Since the multiplier is driven far below its resonance frequency, the phase difference between the pressure inside compared to the pressure outside is nearly zero. It is the small dynamic pressure enhancement inside of the vibromechanical multiplier that drives gas through the regenerator with a velocity that is substantially in phase with the oscillating pressure inside of the bellows created by the oscillation of the linear motor. Inside of the regenerator, the gas undergoes a modified Stirling cycle and moves a net amount of heat from the cold heat exchanger to the ambient heat exchanger. The effectiveness of an acousto-mechanical feedback network [12, 13] compared to a purely acoustic feedback network is shown in Figure 2. This graph shows the magnitude of the three dominant loss mechanisms in an acoustic inertance tube: thermal relaxation of the gas through heat conduction from the gas to the tube wall (blue curve), viscous relaxation of the gas as it oscillates along the tube wall (red curve) and nonlinear entry/exit loss (minor loss) that arises from abrupt changes in cross-section and is partly a function of the abruptness of the cross-sectional area change (green line). The inertance of a tube is expressed as L=ml / A, where l is the length of the tube, A is the cross-sectional area and m is the static density of the working gas. For a constant value of inertance L, the ratio of length and diameter is fixed if the inertance element is assumed to have a circular cross-section of constant diameter.

1 This surround was tested at larger displacement amplitudes than it undergoes in the prototype for over 20 million cycles without failure. The fatigue test was terminated when the testing apparatus failed. While the fatigue limit of the Santoprene has not yet been reached, the two most common failure modes for Santoprene (exposure to ultra-violet light and ozone in the atmosphere) are not present inside the resonator.

10 50

Tube length (cm) 20 30

40

50

60

40 Dissipated Power (W)

Total dissipation Entry/Exit loss Thermal loss Viscous loss

30

20

10

0 2 3 Tube diameter (cm) 4 5

Figure 2: Inertance tube dissipation. This graph shows the losses in an inertance tube that has an inertance of 500 kg/m4. The green dotted curve represents the entry/exit losses and the black curve is the sum of all of the losses in the tube. Notice that while the minimum amount of loss possible is around 13 W, this would require an inertance tube that is almost 1/2 of a meter long and four centimeters in diameter.

A conclusion to draw from this graph is that for an inertance tube with a constant value of inertance (500 kg/m4), the entry/exit loss becomes very large as the tube diameter becomes small. The tube geometry that minimizes the acoustic power dissipation for the ice cream freezer prototype would have to be greater than 4 cm in diameter and over 40 cm long. Using an inertance element this large would have compromised the compactness of the bellows bounce concept and the inertance tube would have still caused over 12 W of acoustic power dissipation. In contrast, the inertial element of the vibromechanical multiplier caused less than 3.5 W of dissipation and is much more compact than an acoustic inertance element. The vibromechanical multiplier cone removes the need to suppress streaming at the expense of a failure mode. In an earlier paper by [14], a tube was used to provide the required inertance for the acoustical phasing network and a latex diaphragm was used to block the toroidal streaming path. The speaker cone provides the inertia and its suspension can easily resist the time-averaged pressure (which imparts a force of less than 1 N on the cone and displaces it less than 0.5 mm from its neutral position) that results from the unidirectional flow of acoustic power through the vibromechanical multiplier.

2.3 Thermal Core


The regenerator is 3 cm thick in the axis of acoustic displacement and is a stack of 290 square stainless steel screens measuring 8.94 cm (3.52) on a side; each screen has 145 wires/inch with a wire diameter of 0.0022. The effective hydraulic radius of the pores produced by the screens is 39 m and the volume porosity is 74% open volume. A required element for a machine that is designed to span a significant temperature is some way to isolate the cold gas from the ambient temperature gas. This is typically part of the job of a displacer in a Stirling machine and is sometimes called a thermal buffer tube in thermoacoustic literature [15]. The thermal buffer in this machine is formed by two windows cut into the thermally insulating plastic2 platforms that hold the regenerator and load heat exchanger.

Ultem 1000, GE Plastics Structured Products, One Plastics Avenue, Pittsfield, MA 01201. http://www.structuredproducts.ge.com

The exhaust (ambient) and load (cold) heat exchangers are identical and consist of a series of parallel flat aluminum tubes that are extruded to include multiple parallel fluid flow channels within the tube3 . The flat tubes have a thickness, b = 1.30 mm, and are spaced by a gap, g = 700m = 2.72 . The ratio of gas cross-sectional area to the total cross-sectional area, Agas / A = g/(b+g). The number of tubes in an array that will completely cover the 80 cm2 regenerator surface is 45. The outer surface area of each tube of length 8.94 cm is 22.5 cm2, so the total surface area of the tube array is 0.10 m2. The heat exchangers have a overall fluid to gas thermal conductance (1/Rth) of approximately 85 W/K; the exhaust heat flow is about 250 W at a typical operating point requiring a temperature difference of 3 K between the regenerator-end and the heat exchanger fluid (about 5% of the total temperature span).

3. PROTOTYPE PERFORMANCE
Table 1 shows the measured performance of the prototype at one operating point and the performance prediction of a detailed DELTAE computer simulation [16] of the prototype. The simulation uses three noted experimentally measured values as input, and the rest of the values in the DELTAE column are the result of the simulators calculations. There is good agreement between the model prediction and the measured quantities. Because of this agreement, the model can be relied upon to provide information about unmeasured quantities like acoustic power dissipation in individual components of the prototype.
Table 1: Prototype performance. This table shows the measured performance of the prototype and a comparison with the performance according to the DELTAE computer simulation. The first three entries in the table are given to DELTAE as inputs, while the rest of the entries in the DELTAE column are results of a convergent DELTAE model. DELTAE predicts poorer performance than is measured mainly due to conservative motor parameters used in the model. The COPc is the COP for an ideal Carnot process.

Quantity Unit Measured DELTAE C input Tload -24.6 0.5 C input Tex 33.9 0.5 Acoustic Pressure kPa 50.81 0.05 input Multiplication Ratio % 8.41 8.48 0.25 Acoustic Power W 134.4 125 3 Electrical Power to Motor W 160.1 147 1.5 Useful Cooling Capacity W 124.6 119 1.2 Exhaust Heat Flow W 296.6 266 2.5 Overall COP 0.78 0.81 0.02 Overall COP/COPc % 18 19 1 The drawing in Figure 3 shows the power flow through the prototype as a Sankey diagram. Starting at the left of Figure 3, 160 W of electrical power is drawn from the amplifier by the linear motor, of which 26 W are dissipated by electrical resistance and eddycurrents in the iron core, mechanical hysteresis in the armature supports and thermoviscous attenuation on the surfaces of the motor. From that electrical input, the motor converts 134 W to acoustic power inside of the bellows. The inner and outer surfaces of the bellows absorb 14 W of this acoustic power from thermoviscous (mostly thermo) attenuation on the surface. Because of the convolutions of the bellows, it has about three times the surface area of a right circular cylinder of the same dimensions, and hence has also three times the acoustic power loss.

Thermalex, Inc., 2758 Gunter Park Drive West, Montgomery, AL 36109; (334) 272-8270. http://www.thermalgroup.com.

Figure 3: Dissipation in the prototype. This Sankey diagram shows the acoustic power flow in the prototype machine for the nominal operating point shown in Table 1. The values shown in this diagram come exclusively from the DELTAE model which has been shown to agree well with performance at this operating point. Values in red highlight the three largest loss mechanisms: the largest is the assumption that all internal dissipation must be pumped by the regenerator which decreases net capacity, secondly, regenerator viscous losses are half of the 2nd Law minimum energy requirement and the surface area of the bellows convolutions is the third largest loss mechanism.

The amount of power delivered to the multiplier from the motor is 120 W and this power is added to 206 W that flows out of the regenerator. The sum, 326 W, is managed by the multiplier with only 5.2 W of thermoviscous (surface) loss and hysteresis loss in the Santoprene flexure that seals the multiplier cone. The 1.6% loss overhead for the multiplier is a real indication of its value in this design. Of the 321 W that are delivered to the thermal core, 10 W is absorbed by the surface of the heat exchangers and 34 W is turned into heat on the surface of the regenerator. This is especially bad, since this waste heat decreases available cooling capacity. Sixty-three watts is the work required by the Second Law of Thermodynamics to move 190 W of heat from the cold temperature (-24.6 C) to the exhaust temperature (29.0 C). In other words, even a perfect regenerator operating at these temperatures would dissipate 63 W of energy and be operating at limiting Carnot COP. Although the Sankey diagram of Figure 3 shows only the flow of acoustic power, the arrows at the load heat exchanger show the various heat loads that the machine moves up the temperature gradient. The potential cooling capacity at this operating point is 190 W. This value is reduced to 124.6 W by two nuisance loads: the 57 W of internal surface dissipation described in the above paragraph that must be pumped and about 8 W of heat flow from the

room to the cold side of the machine. The estimate of heat leak was made by calculating the thermal resistance of the plastic platforms and the bolts used to close the pressure vessel. At last, the flow of acoustic energy that leaves the regenerator is fed back to add to the energy generated by the motor. The surface of the thermal buffer tube formed by the windows in the regenerator platform and the flow straightening screens dissipate less than 8 W of acoustic power before the feedback path converges with the power supplied by the linear motor.

3. PROTOTYPE VOLUME
The table below shows the power density of various refrigeration machines, adjusted for their temperature lift as follows: & T Q &= V T Q load

(1)

where V is the volume displaced by the entire machine, & is the cooling capacity, T is the Q temperature lift and Tload is the temperature of the load from which heat is being removed by the machine.
Table 2: Power density comparison. This table shows the normalized power density of three electrically driven thermoacoustic refrigerators. TRITON and SETAC were both dual-stack machines driven at the lowest acoustic mode of the resonator. The volume figure in the table is the volume of the envelope of the resonator, which is significantly larger than the space occupied by gas in the machine. Since these were research machines, their envelope volume was not optimally minimized as would be done for a commercial product. Although not fully comparable, a volume figure for a hermetic refrigerant vapour compressor is added.

Machine TRITON SETAC Prototype Embraco Mod: EMI45HER

Capacity Volume T 10 kW 294 W 125 W 126 W 1000 l 47 l 24 l 7l 17 K 37 K 60 K 70 K

Tload 281 K 276 K 248 K 248 K

& Q 0.6 W/l 0.6 W/l 1.3 W/l 5.1 W/l

As seen in Table 2, the Prototype is a factor of two more compact than other electrically driven thermoacoustic research prototypes built to date. The power density of the commercially available refrigerant vapour compressor is about a factor of four better than the prototype, according to the Embraco4 catalog posted on their web site. However, this estimate does not take into consideration that the prototype has over 25% more volume than would be required for it to function; this extra space was included to instrument the prototype and allow for easy disassembly and reassembly. Therefore, further volume optimization is possible. Initial trials show that it is possible to locate the prototype in its current dimensions into a standard 200 liter ice cream sales cabinet.

Embraco, Rua Rui Barbosa 1020 Cep: 89219-901, Joinville Brazil http://www.embraco.com.br

4. CONCLUSIONS
The prototype thermoacoustic refrigerator as described in this publication, meets prescribed specifications of efficiency and size. The prototype is therefore able to maintain ice cream at the proper serving temperature in a standard 200-liter sales/storage cabinet. As a result of good agreement between measurable quantities and a computer simulation of the performance of the prototype, the model can be mined for the details of the performance of each subsystem. This analysis results in a clear vision for further improvement of the prototypes performance. Along with gains in efficiency and size, new regenerators and flexure seals should be created that are cheaper than the screens and metal bellows used today. The opportunities that still lie ahead for the technology of thermoacoustics are very exciting, both scientifically and commercially.

5. ACKNOWLEDGEMENTS AND REFERENCES


This work was generously supported by Unilever, Ben & Jerrys Homemade, Inc. and the Penn State Applied Research Laboratory.

1. 2. 3. 4. 5. 6. 7. 8. 9. 10.

11. 12. 13. 14. 15. 16.

Backhaus, S., and Swift, G. W., Nature, 399, 335338 (1999). Garrett, S. L., Am. J. Phys., 72, pp. 1117 (2004). Garrett, S. L., and Backhaus, S., American Scientist, 88 (2000). Garrett, S. L., Adeff, J. A., and Hofler, T. J., J. Thermophys. Heat Transfer, 7, pp. 595 599 (1993). Garrett, S. L., High-power thermoacoustic refrigerator (1997), U. S. Patent No. 5,647,216. Poese, M. E., and Garrett, S. L., J. Acoust. Soc. Am., 107, 24802486 (2000). Backhaus, S., and Swift, G. W., J. Acoust. Soc. Am., 107, 31483166 (2000). Gedeon, D., DC Gas Flows in Stirling and Pulse-Tube cryocoolers, in Cryocoolers 9, edited by R. G. Ross, Plenum Press, New York, 1997, pp. 385392. Poese, M. E., Smith, R. W. M., Wakeland, R. S., and Garrett, S. L., Compliant enclosure for thermoacoustic device (2003), U. S. Patent App. No. 2003/0192323. Smith, R. W. M., High Efficiency Two Kilowatt Acoustic Source for a Thermoacoustic Refrigerator, Masters thesis, The Pennsylvania State University (Dec. 2001), also ARL Technical Report No. TR-01-001. Smith, R. W. M., Poese, M. E., Wakeland, R. S., and Garrett, S. L., Thermoacoustic device (2003), U. S. Patent No. 6,725,670. Swift, G. W., Backhaus, S. N., and Gardner, D. L., Traveling-wave device with mass flux suppression (1999a), U. S. Patent No. 6,032,464. de Blok, C., Thermo-acoustic system (1998), International PCT Pub. No. WO 99/20957 and U. S. Patent No. 6,314,740. Swift, G. W., Gardner, D. L., and Backhaus, S., J. Acoust. Soc. Am., 105, 711724 (1999b). Swift, G., Thermoacoustics: A unifying perspective for some engines and refrigerators, Acoustical Society of America ( http://asa.aip.org) 2002. Ward, W. C., and Swift, G. W., J. Acoust. Soc. Am., 95, 36713672 (1994).

You might also like