You are on page 1of 13

Development of MEMS based pyroelectric thermal energy harvesters

Scott R. Hunter
ABSTRACT
The efficient conversion of waste thermal energy into electrical energy is of considerable interest due to the huge sources of low-grade thermal energy available in technologically advanced societies. Our group at the Oak Ridge National Laboratory (ORNL) is developing a new type of high efficiency thermal waste heat energy converter that can be used to actively cool electronic devices, concentrated photovoltaic solar cells, computers and large waste heat producing systems, while generating electricity that can be used to power remote monitoring sensor systems, or recycled to provide electrical power. The energy harvester is a temperature cycled pyroelectric thermal-to-electrical energy harvester that can be used to generate electrical energy from thermal waste streams with temperature gradients of only a few degrees. The approach uses a resonantly driven pyroelectric capacitive bimorph cantilever structure that potentially has energy conversion efficiencies several times those of any previously demonstrated pyroelectric or thermoelectric thermal energy harvesters. The goals of this effort are to demonstrate the feasibility of fabricating high conversion efficiency MEMS based pyroelectric energy converters that can be fabricated into scalable arrays using well known microscale fabrication techniques and materials. These fabrication efforts are supported by detailed modeling studies of the pyroelectric energy converter structures to demonstrate the energy conversion efficiencies and electrical energy generation capabilities of these energy converters. This paper reports on the modeling, fabrication and testing of test structures and single element devices that demonstrate the potential of this technology for the development of high efficiency thermal-to-electrical energy harvesters.

Key Words:

Energy harvesting, pyroelectric, bimorph cantilever, MEMS, surface micromachining

1.
1.1. Background

INTRODUCTION

Industry, world-wide, discharges over 1001012 joules (TJ) annually of low-grade waste heat (10C to 250C) from electric power stations, pulp and paper mills, steel and other metal foundries, glass manufacturers and petrochemical plants1. In the U.S. around 55% of the energy generated from all sources in 2009 was lost as waste heat2. A technology to recover or convert this low-grade waste heat to usable electricity could save industrial sectors tens of billions of dollars annually, through increased process efficiencies and reduced fuel costs, while substantially reducing greenhouse gas emissions. Other opportunities also exist for active cooling and electrical power generation for sensor systems on much smaller scales, such as on-chip active heat sinks, in standalone computer systems and computer data processing centers.

The useful work content of all thermal engines is thermodynamically limited by the Carnot efficiency, Carnot:

Carnot = 1 TL / TH

(1)

where TH is the temperature of the heat source and TL is the temperature of the heat sink. Thermal energy gradient power generators convert heat (Qin) into electrical energy (Wout) with efficiency:
(2)

where WE is the generated electrical energy, WP is the energy lost in the temperature cycle, C is the heat capacity of the pyroelectric device, QInt are the intrinsic heat losses in the thermal cycle and QLeak are the heat leakages between the hot and cold sources. Presently contemplated thermal to electrical energy conversion techniques (thermoelectric and pyroelectric) all suffer from low energy conversion efficiencies, limited partly by the Carnot efficiency, but also by the inherent limitations of the conversion technologies themselves. Pyroelectric converters remain relatively unexplored, as early attempts to model and fabricate converters based on pyroelectric operating principles gave uneconomically low conversion efficiencies (0.1-2%)3,4. Other modeling studies were much more encouraging however, with overall predicted energy efficiencies of 10-15%5 and 20-40%6-8 with Carnot efficiencies in the range 50-80%610 or higher11-19. In contrast, thermoelectric generators have maximum Carnot efficiencies around 14-17%9,10,20 and overall efficiencies around 5%21. Pyroelectric energy generators rely on the property that the spontaneous polarization (and hence dielectric constant) of certain materials is temperature dependent. Cycling the materials temperature induces an alternating current in an external circuit when the pyroelectric material is made the dielectric in a capacitor. This property is shown schematically in Figure 1, where the intrinsic dipole moment of the pyroelectric material is made part of a capacitor and an ammeter connected between the two capacitor electrodes 22. At constant temperature, no current flows in the circuit. When the capacitor temperature is increased, the polarization P S decreases, effectively reducing the capacitors dielectric constant, and causing a current to flow in the external circuit to compensate for the decrease in the bound charge in the capacitor22. This property can be used to generate electricity where the electrical current and energy conversion efficiency depends on the rate of change, and on the magnitude of the

Figure 1. A pyroelectric material used as the dielectric in a capacitor. As the temperature of the capacitor is increased, a current flows in the external circuit to compensate for the change in bound charge at the edges of the crystal22.

temperature change in the capacitor.

1.2

Ericson Thermal Energy Conversion Cycle

A quasi-isothermal cycle, such as the one shown in Figure 1, is very inefficient and, as a result, produces very little power. However, if the Ericson thermal energy cycle is used to extract electrical energy from the thermal gradient, much higher thermal-to-electrical conversion efficiencies and output powers are achievable 11-19. The cycle is shown in Figure 2 and works by allowing large temperature swings across the pyroelectric capacitor while applying alternating voltages on the capacitor electrodes. An energy scavenging, pyroelectric capacitor circuit utilizing this temperature cycle is shown in Figure 3. The cycle starts at (a) in Figure 2 with the pyroelectric capacitor at low temperature T L and the ferroelectric capacitor charged at high voltage V 2. As the temperature increases to T H at a constant applied voltage (b), charge is forced to flow in the external circuit charging the storage capacitor (Figure 3). The applied voltage is then reduced to V1 at (c) and the temperature of the pyroelectric capacitor decreased to T L again (d), producing another, opposite sign, current flow in the external circuit. Other thermal cycles include Rankin and Stirling cycles and are used in steam power plants, internal combustion engines and refrigerators.

S w i t c h in g a n d S u p p ly C o n t r o l C i r c u i t r y V o l t a g e , C u r r e n t a n d T e m p e r a t u r e S e n s in g S ig n a l s P y r o e le c t r i c C a p a c i to r A r ra y

V
V o l ta g e S u p p ly

C u r re n t R e c t i f ic a t io n C ir c u it

S to r a g e C a p a c i to r

Load

Figure 2. The temperature and voltage cycle used to generate electrical energy in the pyroelectric circuit shown in Figure 3.

Figure 3. A schematic showing an electrical circuit that can be used to generate electrical power from an array of pyroelectric capacitors.

The pyroelectric current Ip produced during the cycle shown in Figure 2 is:

I p = Af

dPS dT = Af p dt dt

(3)

where Af is the surface area of the pyroelectric thin film capacitor, PS (C/m2) is the pyroelectric thin film polarization (Figure 1), T is the pyroelectric capacitor temperature and p is the pyroelectric coefficient in C/m 2K. The net output power Np from the pyroelectric capacitor is:

N p = Vappl I p = Vappl pAf

dT dt

(4)

where Vappl is the external applied voltage across the pyroelectric capacitor (Figure 3). The cumulative pyroelectric conversion output work Wout from the cycle is as follows: (5)

Wout =

appl

dq =

N dt = V
p

appl

pAf

dT dt dt

Equation 5 is shown schematically in Figure 2 where Wout is the integral over the area within the figure: the greater the change in applied voltage across the pyroelectric capacitor and the wider the temperature swing, the larger the amount of heat energy converted to useful electrical energy. Equations 3 and 5 also show that the amount of current and electrical energy generated by this circuit is dependent on the magnitude of the pyroelectric coefficient p, the size of the capacitor (plate area A), and very importantly, on the rate of change in the temperature across the pyroelectric capacitor. Hence the faster the temperature can be cycled back and forth across the device, the more efficient the energy conversion process is and the greater the amount of electrical energy generated.

1.3

Pyroelectric Energy Harvesters

Prior attempts to use this technique to generate electricity have suffered from low energy conversion efficiencies due to the low operating frequencies (< 1 Hz), large power requirements to generate significant temperature cycles (Wp in Equation 2), large thermal mass capacitor systems with relatively low breakdown strengths (i.e. low voltage differences, V2-V1) and low thermal conductivities leading to low T/t, and hence low Q/T6-19,23-39. The MEMS based pyroelectric power generator outlined in this paper operates at much higher frequencies (10s of Hz to 100s Hz), using thin film structures with low thermal masses and comparatively high dielectric strengths, and high thermal conductivities (giving fast T/t and hence large Q/T). The innovative use of the heat source to power the temperature cycling through the converter using bimaterial heat sensitive structures, and use of resonantly driven cantilever motion to rapidly move the converter through the temperature cycle leads to high efficiency operation (i.e. WP 0 in Equation 2). Encapsulating the generator in a partially evacuated enclosure also minimizes heat losses through gas convection and conduction processes (i.e. QLeak 0 in

Equation 2). Consequently, expected conversion efficiencies of a fully optimized converter will be as high as 8090% of the Carnot limit. The actual energy conversion efficiency for any thermal energy recovery device depends on the temperature difference between the hot and cold sources. For temperature differences in the range 10-20 0C, such as those found in cooling systems for supercomputer and data processing centers, maximum overall efficiencies in the range 3-7% are achievable. With higher temperature differences in the range 100-300 0C, such as those for computer microprocessors, internal combustion engines and steam power plants, overall efficiencies in the 20-40% range are achievable. These values are several times larger than those obtainable with any of the other competing thermal-to-electrical scavenging techniques. In this paper, we outline the design, modeling and preliminary testing we have performed on MEMS based pyroelectric capacitive structures, to show the principle of operation and potential for scale up of this technology to larger more efficient energy harvesters.

2.
2.1

MEMS BASED PYROELECTRIC THERMAL ENERGY HARVESTER


Energy Harvester Concept

Figure 4. Schematic of a pyroelectric energy harvester device, consisting of a bimaterial cantilever structure, which alternately contacts the hot and cold surfaces, generating an electrical current in the pyroelectric capacitor.

Figure 5. Schematic showing the details of the construction of the capacitive cantilever thermal energy converter and split anchor structures.

The concept for an individual thermal to electrical energy generator element consisting of a cantilevered pyroelectric capacitor structure is shown schematically in Figure 4. Each energy converter structure is a few hundred m to several mm in length and width. The cantilever structure is shown in detail in Figure 5, and is fabricated with two metal films, which act as the electrodes of a capacitor, and a pyroelectric material (e.g. copolymer of poly-vinylidene fluoride (P(VDF-TrFE)) which acts as the dielectric between the metal electrodes. Two additional small proof masses may be located at the ends of the cantilever to increase the thermal mass of the structure and to make good thermal contacts with the hot and cold surfaces. The cantilever can be anchored to either the hot or cold surfaces. A split anchor also provides the capacitor electrical contacts to the external charge extraction and control circuitry (Figure 3). These millimeter scale converters can readily be scaled up to much larger devices using 2D arrays of individual converter elements. Arrays up to 10 6 converters can be fabricated, and these arrays themselves can be stacked to harvest electrical energy from much larger thermal waste energy sources.

2.2

Energy Harvester Operation

A side on schematic view of the harvester structure is given in Figure 6, showing the approximate relative thicknesses of the materials making up the bimorph cantilevered structure. The P(VDF-TrFE) pyroelectric layer has a large coefficient of thermal expansion (TCE) while the thicker lower metal layer shown in Figure 6 is

composed of a low CTE metal, such as Ti. The upper thin metal layer is sufficiently thick (typically 10-50 nm) that it forms a continuous metal film over the P(VDF-TrFE) dielectric layer, but not so thick that it contributes to the bimorph bending of the cantilever. The bimorph metal and P(VDF-TrFE) layers are typically 2-10 m thick. The harvester operation is shown schematically in Figure 7. The cantilever structure initially heats through the anchor, causing the cantilever to bend towards the lower cold heat sinked surface. On contacting the cold surface, the structure rapidly loses heat, and bends towards the upper hot surface. On contacting the upper surface, it then rapidly heats and bends from the upper surface and again makes contact with the lower surface. This process is repeated indefinitely. Good thermal contact with the hot and Figure 6. Side view of the pyroelectric bimorph cantilever cold surfaces is essential to transfer substantial amounts structure showing the cantilevered capacitor thin film layers. of thermal energy (and thus to raise the cantilever temperature substantially) to the pyroelectric capacitor on the cantilever. Heat transfer is facilitated by small stiction forces between the cantilever proof masses and the hot and cold surfaces, which will counteract the bimorph mechanical force pulling the cantilever structure away from the surfaces once contact has been made and the temperature increased or decreased in the structure. Cantilever-surface contact time can also be controlled by Figure 7. Schematic showing the operation of the bimorph cantilevered alternately applying an electric potential pyroelectric capacitor structure alternately contacting both the hot and cold (10-80V) between the proof masses and surfaces. the adjacent surface. The longer the cantilever is in good thermal contact with the hot or cold surfaces, the more energy that is transferred to the pyroelectric capacitor and the hotter the bimorph structure, the more mechanical force that is imparted to the cantilever to drive it toward the opposite surface once the restraining forces are removed (i.e. the mechanical force overcomes the stiction force and the electrostatic force once the electrical potential is removed). An approximate timing diagram for the operation of this electro-mechanical circuit is shown in Figure 8. The upper trace shows the temperature in the pyroelectric capacitor as the cantilever alternately contacts first the lower cold surface, then the upper hot surface. The thermal response time is dependent on the thermal contact resistance between the proof masses and the hot and cold surfaces, and heat capacity of the cantilever and capacitor structures. The next lower trace shows the change in pyroelectric capacitance in response to the temperature change. Also shown in Figure 8 is the timing of the applied voltages across the capacitor to implement the Ericson thermal energy extraction cycle. The letters

Figure 8. An approximate timing diagram showing the change on pyroelectric capacitor temperature and capacitance as the cantilever alternately contacts the hot and cold surfaces. Also shown is the timing of the applied voltage form the temperature cycle in Figure 2 and the resulting expected current extracted from the circuit shown in Figure 3.

correspond to the points on the temperature-voltage cycle diagram shown in Figure 2. The field is switched just prior to the cantilever contacting the hot or cold surfaces. The final curve in Figure 8 shows the expected rectified current extracted from the electrical circuit shown in Figure 3. The initial sharp pulse occurs due the field switching across the pyroelectric capacitor while the second shows the current extracted from the capacitor due to the temperature changes cross the capacitor while the field is applied. The power used in voltage switching across the capacitor is not lost but also goes into charging the storage capacitor shown in Figure 3.

3.

ENERGY HARVESTER OPERATION MODELING

We have performed extensive modeling of the mechanical and heat transfer behavior of a bimaterial cantilever placed in proximity of a heat source using finite element analysis (FEA) techniques in the multiphysics COMSOL modeling package. Our models combined solid mechanics, heat transfer physics and electrostatics selected from the standard set of COMSOL multiphysics modules. We used a 2D model in order to maintain a reasonably low number of degrees of freedom to be solved in the model while defining mesh sizes sufficiently small to accurately reflect the smallest characteristic features of our structures. Typically, our models consist of up to 1000 mesh elements and up to 10000 degrees of freedom solved. The meshing around the contact point between the cantilever and the heat source is particularly important. The meshing used in our modeling and its quality is shown in figure 9.

Figure 9. Illustration of the

The main geometrical parameters of the modeled structures, along with model geometry, mesh and the material properties used in our models, are given in Table 1. We used a time mesh quality in the vicinity of the thermal contact between the dependent solver in order to simulate both cantilever deformations and the cantilever (bottom) and heat temperature distribution in our system as a function of time. As an approximation source (top). The color scale of idealized experimental conditions, we used thermal insulation boundary from blue to dark red conditions on all cantilever surfaces except its anchoring surface, and a convex correspond to the mesh quality surface that provided a contact between the cantilever top surface and the heat 0 to 1, respectively. source. The constant temperature boundary condition was used on the side and top surfaces of the heat source as well as the anchoring surface of the cantilever. We constructed our model with convex surfaces in the vicinity of the anticipated contact between the cantilever and the heat source, in order to mimic individual microscale asperities present in a typical solid-to-solid contact due to surface roughness and undulations. The presence of convex surfaces also allowed us to eliminate any ambiguity and shift in the x- coordinate of the contact point that would otherwise have taken place between two flat surfaces due to their deformation upon heating.

Table 1. Geometrical and Materials Properties of the Modeled Structures


Material Youngs modulus (Pa) 70 73 Poissons ratio 0.33 0.17 Themal conductivity (W/mK) 160 1.3 Coefficient of thermal expansion (ppm/K) 23 0.55 Density (kg/m3) 2700 2200 Heat capacity (J/kg K) 900 703

Al SiO2

The intermittent thermal contact between the heat source and the top cantilever surface was modeled by defining the two convex surfaces as an identity pair with heat continuity boundary conditions. While this represents a somewhat crude approximation of the highly complex behavior of the actual thermal contacts between two solids surfaces, we expect it to describe our experimental system reasonably well by empirically adjusting the thermal conductivity of the region in vicinity of the thermal contact (Figure 10). In our simulations, we varied thermal conductivity of this region in the range of 0.3 to 30 W/mK. The other variable parameters of our model included cantilever length, L = 1 mm and L = 5mm, temperature difference between the cantilever and the heat source, T (10, 50 and 150 K) and an optional force applied to the cantilever tip in the vertical direction ,either up

Cantilever anchor/clamp

Contact area with variable thermal conductivity, k cont

Heat Source

Aluminum SiO2

or down. The latter was used to mimic the effect of electrostatic interactions that can be introduced into our experimental system in order to control the self-oscillatory behavior. Figure 11 shows typical modeling results obtained from the COMSOL simulation. In this figure the temperatures of the cantilever and the heat source, the cantilever deflection and the direction of the heat flux in the cantilever are shown at the time of contact between the cantilever and the heat source (left image) and approximately 1 msec after the cantilever has separated

Cantilever

Figure 10. Model geometry with a color scale corresponding to thermal conductivity of the components used in our model.

Figure 11. Two images showing the moment of contact between the cantilever and the heated source and approximately 1 ms after the cantilever has separated from the source. The images show the cantilever and heat source temperatures, the cantilever tip displacement and the arrows show the direction of the heat flux down the cantilever.

from the heat source (right image). Figure 12 shows the cantilever tip displacement and the temperature of the cantilever at the point of contact with the heat source as the cantilever repeatably makes contact with heat source. Figure 13 shows the change in cantilever temperature at the contact point and at two points along the cantilever toward the base for the same simulation conditions. The initially large temperature fluctuations near the cantilever tip rapidly decrease in magnitude away from the point of contact with the heat source. The cantilever is oscillating at around 50 Hz under these conditions.

Figure 12. Cantilever tip displacement (blue) and temperature (red) analyzed at the cantilever tip contact point as a function of time for T=150K, kcont=3 W/mK and L=1mm.

Figure 13. Temperature changes analyzed at the cantilever tip contact point and at points 200 m and 500 m from the tip as a function of time for T=150K, kcont=3 W/mK and L=1mm.

Figures 14 and 15 summarize the heat flow across the cantilever-heat source contact point as functions of T before contact and the thermal conductivity across the contact point. Electrical power generation from the device is critically dependent on the heat flow through the cantilever structure, which in turn, is very dependent on the temperature difference and thermal conductivity across the contact point. The heat flux across the contact is also dependent on the time that the two surfaces are in contact. If the cantilever barely touches the heat source before bouncing off, very little heat will be transferred to the cantilever, resulting in minimal temperature fluctuations across the pyroelectric capacitor located on the cantilever.

Figure 14. Heat flux through the self-oscillating cantilever calculated per 1mm of the cantilever width as a function of T.

Figure 15. Heat flux through the self-oscillating cantilever calculated per 1mm of the cantilever width as a function of the contact region thermal conductivity used in the model.

Normal microsystems stiction forces will tend to prolong the contact time until the temperature increase, and the resultant bimorph bending force in the cantilever, is sufficient to overcome the restraining forces on the cantilever. These forces result in unpredictable contact times as was shown in the experimental measurements on bimorph cantilever structures described in the next section. Contact times can be enhanced and controlled by using electrostatic forces to hold the cantilever and heat source in contact for fixed time periods. These forces have been simulated by applying a static force to the cantilever to understand their influence on the cantilever motion and change in temperature. The results from one of these simulations is shown in Figure 16 along with the situation where no additional forces are present. Application of a static force has the effect of initially producing a large increase in T, but subsequent oscillations are much more rapid, resulting in much smaller temperature changes. Applying the electrostatic force only while the cantilever is in contact with the heat source will improve contact and heat transfer. These simulations demonstrate the complexity of the task required to accurately simulate and predict the operation and efficiency of the heat transfer process. Similar calculations are required to understand operation of the device once the pyroelectric energy conversion capacitors are incorporated in the cantilever structures. We intend to perform these calculations in subsequent studies.

Figure 16. Cantilever tip displacement (dotted lines) plotted together with temperature of the cantilever tip (solid lines) as a function of time for T=150K, kcont=3 W/mK, L=1mm and two different values of force applied to the tip in the direction of the heat source. F= 0 N (red lines); F=2x 10-4 N (blue lines).

4. 4.1 Device Characterization

EXPERIMENTAL DESIGN VERIFICATION

A vacuum test chamber and data collection system was fabricated to temperature cycle the bimorph test cantilever structures and to characterize cantilever thermal heat transfer structures and the pyroelectric and electrical current generating properties of pyroelectric AlN and P(VDF-TrFE) capacitive structures. The chamber can be pumped to vacuum pressures in the range 20-50 mTorr, which is sufficient to eliminate heat loss by gas conduction and convection in the vacuum chamber. The test setup is shown in Figure 17, and, in addition to the vacuum chamber, consists of a Labview controlled temperature controller and thermoelectric (TE) cooler, a firewire video camera used to obtain video images of the moving cantilever structures (examples are shown in Figure 19), and a precision three dimensional translation stage that can be used to accurately position the cantilever structures between the heat source and sink. A cantilever position sensing laser and position sensitive detector (PSD), which can be used to monitor the motion of the tip of the cantilever structures as a function of heat source

temperature40 is also shown in the left hand image of Figure 17. The right hand image of Figure17 is an enlarged view of the cantilever setup inside the chamber, showing the bimorph cantilever and support structure (silver mirror) mounted on the TE cooler (gold plated square structure), with the hot wire source located near tip of the cantilever (the wire support post is reflected in the silver cantilever support structure). Thermocouple sensors are mounted on the TE cooler to rapidly sense changes in the cantilever temperature. Figure18 shows an image frame taken by a video camera showing the bent cantilever located top of the TE cooler with the hot wire source located just above the tip of the cantilever. The two left hand images in Figure 18

Figure 17. Vacuum test chamber and optical diagnostics used to characterize the performance of the resonating cantilever structures (left) and a magnified image of the cantilever, TE cooler and heated wire source (right).

Figure 18. Video (right) and IR (left) images of the hot wire source and the cantilever making contact with the heat source. The IR images were used to estimate the temperature of the heated wire source.

show infrared images of the cold and heated wire. This imagery allows us to determine the wire temperature and estimate heat flow to the bimorph cantilever. Figure 19 depicts a series of frames from a video showing a self resonating bimorph cantilever moving in response to changes in the substrate temperature. The tip of the cantilever is moving approximately 1 Figure 19. Several frame shots showing one oscillation cycle of a mm in response to substrate temperature resonantly heated cantilever structure. The cantilever resonated with 0 changes of +/- 5 C. This is an extremely frequencies around 30 Hz. large range of motion for this device and will be more than enough to allow thermal cycling through the device when the cantilever is made to contact both the heat source and the cool heat sink. The next step will be to show resonant cantilever motion by having the heated cantilever make contact with the cold copper heat sink.

4.2

Temperature Cycled Pyroelectric Capacitors

Figure 20. An array of fabricated mm. sized P(VDF-TrFE) test capacitors used to estimate the thermal to electrical power conversion efficiency.

Arrays of pyroelectric test capacitors were fabricated using two types of pyroelectric materials, AlN and polyvinylidene difluoride-trifluoroethylene copolymer (PVDF-TrFE or copolymer), to explore the issues involved in the fabrication and operation of these temperature cycled capacitive structures, and their integration into resonating cantilevers to form thermal energy conversion devices. This copolymer material has been used in many of the previous attempts to fabricate pyroelectric thermally cycled energy converters due to its relatively high pyroelectric coefficient and dielectric strength, and ease of fabrication using standard wet chemical and wafer processing equipment and techniques 11-19,23-36.

An example of one of these arrays of fabricated capacitors is shown in Figure 20. These capacitors were made by sputtering 3 mm2 Al electrodes onto each side of a 25 m thick, electrically poled P(VDF-TrFE) copolymer film. Electrical leads were attached to the capacitor electrodes using silver based electrically conductive paste or removable probes when individual capacitive elements in an array were studied (Figure 20). The measured capacitances for several of these thin film pyroelectric capacitors ranged from 10 pF up to 1 nF. A 10 pF capacitor was mounted on the temperature cycled TE cooler shown in Figure 17, and connected to the data acquisition system and temperature cycled over a range of maximum and minimum temperatures. An example of one these temperature cycling measurements, along with the resultant current generated in the capacitor, is shown in Figure 21. A summary of the peak current as a function of the peak heating rate from these measurements is shown in Table 2. As expected from Eq. 3, the measured current is a function of the rate of change in Figure 21. Temperature cycled P(VDF-TrFE) test capacitors using the TE cooler shown in Figure to cycle temperature across the pyroelectric capacitor. Similar the temperature between 100C and 600C. Note the current measurements with a flash heating lamp allowed us to spikes when the rate of change in the temperature is obtain very high rates of temperature change in these greatest. capacitors, giving T/t 5,000 K/s. These extremely high rates gave peak currents Ip 750nA.

Table 2. Measured peak pyroelectric capacitor currents as a function of peak rates of change in temperature cycling Peak Heating Rate Peak Current
0

1 .36 .5 0 .25 .7

4 .4 0 .2

7 1.3 1 .8

1 3.4 1 .2

1 2

C/s n A

These measurements show that high rates of temperature change across the capacitive structures are key to obtaining high rates of heat to electrical energy conversion. Incorporating these pyroelectric capacitors in the resonating cantilever structures, and operating these devices at frequencies up to 100 Hz with temperatures gradients of 20-500C or higher, will allow us to operate the energy harvester at these extremely high temperature cycling rates. Further, as outlined in Section 1, the quasi-isothermal approach used to obtain these results is very inefficient in generating electrical power. By temperature and voltage cycling these capacitors using the Ericson thermal engine cycle approach, much higher conversion efficiencies are achievable 9-19. We intend to implement this energy conversion cycle during future work on this project.

5.

DISCUSSIONS AND CONCLUSIONS

We have developed a new innovative concept for the conversion of low temperature thermal waste heat into electrical energy using electromechanically resonating MEMS based microstructures, which convert steady state thermal energy gradients into temporally varying energy gradients across these millimeter sized structures.

10

Pyroelectric materials can be incorporated as dielectrics in capacitive elements within these MEMS structures, and used to efficiently generate electrical energy from the resulting time varying energy gradients. These millimeter sized energy harvester structures can be fabricated in 2D arrays and scaled up in size allowing these devices to generate sizable amounts of electrical power for many applications. The design, modeling and experimental work outlined in this report describes the initial studies we have performed to demonstrate the utility of this technique. Although these studies are in their initial stages, we have for the first time, modeled and demonstrated a resonantly activated cantilever structures vibrating at frequencies of several tens of cycles per second, operating entirely within a steady state temperature gradient. These operating frequencies (and resultant rates of temperature change in the cantilever structures) are several orders of magnitude greater than have previously been achieved in pyroelectric energy harvesting devices. Secondly, we have performed extensive temperature cycling studies on fabricated thin film pyroelectric capacitive structures and demonstrated the huge thermal to electrical energy conversion advantage that can be achieved by operating these devices at high frequencies. An additional advantage of this technique is that the energy harvesters can be scaled to arbitrarily large 2D arrays using standard microscale and semiconductor fabrication techniques for use in many applications. These studies indicate we can achieve our goal of developing a compact, efficient thermal to electrical energy pyroelectric harvester using electromechanical resonating structures.

ACKNOWLEDGEMENTS
Research sponsored by the Laboratory Directed Research and Development program at the Oak Ridge National Laboratory, managed and operated by UT-Battelle, LLC, for the U. S. Department of Energy under Contract No. DE-AC05-00OR22725.

REFERENCES
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] Ikura, M., Conversion of low-grade waste heat to electricity, CANMET Energy Technology Centre, http://canmetenergy-canmetenergie.nrcan-rncan.gc.ca/eng/publications.html?2008-56 Lawrence Livermore National Laboratory, Estimated U.S. energy use in 2009, https://flowcharts.llnl.gov/ Fattuzzo, E., H. Kiess and R. Nitshe, Theoretical efficiency of pyroelectric power converters, J. Appl. Phys. 37, 510-516 (1966). van der Ziel, A., Solar power generation with the pyroelectric effect, J. Appl. Phys. 45, 4128 (1974). Gonzalo, A., Ferroelectric materials as energy converters, Ferroelectrics 11, 423-430 (1976). Drummond, A.D., Dielectric power conversion, Abstracts 10 th Intersociety Energy Conversion Engineering Conf., Newark, DE, 950 (1975). Drummond, A.D., Dielectric thermal power converter, U.S. Patent 4,220,906 (1980). Drummond, A.D., Demonstration of a high power density electrocaloric heat engine, Ferroelectrics 27, 213-218 (1980). Sebald, G., E. Lefeuvre and D. Guyomar, Pyroelectric energy conversion: optimization principles, IEEE Trans. Ultrasonics, Ferroelectrics, & Frequency Control 55, 538-551 (2008). Guyomar, D., G. Sebald, E. Lefeuvre and A. Khodayari, Toward energy harvesting using pyroelectric material, J. Intell. Mat. Sys. Struct. 20, 265-271 (2009). Olsen, R.B., J. M. Briscoe, D. A. Bruno and W. F. Butler, A pyroelectric converter which employs regeneration, Ferroelectrics 38, 975-978 (1981). Olsen, R.B., Ferroelectric conversion of heat to electrical energy a demonstration, J. Energy 6, 91 (1982). Olsen R.B. and D. D. Brown, High efficiency direct conversion of heat to electrical energy-related pyroelectric measurements, Ferroelectrics 40, 17-27 (1982). Olsen, R.B. and D. Evans, Pyroelectric energy conversion: hysteresis loss and temperature sensitivity of a ferroelectric material, J. Appl. Phys. 54, 5941-5944 (1983).

11

[15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39]

Olsen, R.B., J. C. Hicks, M. G. Broadhurst and G. T. Davis, Temperature dependent ferroelectric hysteresis study in polyvinylidene fluoride, Appl. Phys. Lett. 43, 127-129 (1983). Olsen, R.B., D. A. Bruno, J. M. Briscoe and J. Dullea, Cascaded pyroelectric energy converter, Ferroelectrics 59, 205-219 (1984). Olsen, R.B., D. A. Bruno and J. M. Briscoe, Pyroelectric conversion cycles, J. Appl. Phys. 58, 47094716 (1985). Olsen, R.B., D. A. Bruno, J. M. Briscoe and E. W. Jacobs, Pyroelectric conversion cycle of vinylidenefluoride-trifluoroethylene copolymer, J. Appl. Phys. 57, 5036-5042 (1985). Olsen, R.B. Pyroeletric energy converter element comprising vinylidene fluoride-trifluoroethylene copolymer, U.S. Patent 4,620,262 (1986). Bierschenk, J., Optimized thermoelectrics for energy harvesting applications, in Energy Harvesting Technologies, (S. Priya and D. J. Inman, Eds.) Springer, New York, Chpt. 12, 337-351 (2009). Snyder, J.G., Thermoelectric energy harvesting, in Energy Harvesting Technologies, (S. Priya and D. J. Inman, Eds.) Springer, New York, Chpt. 11, 323-336 (2009). Lang, S., Pyroelectricity: From ancient curiosity to modern imaging tool, Phys. Today 58, 31-36 (2005). Kouchachvili, L. and M. Ikura, Pyroelectric conversion Effect of P(VDF-TrFE) preconditioning on power conversion, J. Electrostatics 65, 182-188 (2007). Kouchachvili, L. and M. Ikura, Improving the efficiency of pyroelectric conversion, I. J. Energy Res . 32, 328-335 (2008). Kouchachvili, L. and M. Ikura, High performance P(VDF-TrFE) copolymer for pyroelectric conversion, U.S. Patent 7,323,506 (2008). Ikura, M., Pyroelectric conversion system, U.S. Patent 6,528,898 (2003). Ikura, M., Conversion of low-grade heat to electricity using pyroelectric copolymer, Ferroelectrics 267, 403-408 (2002). Capineri, L., L. Masotti, V. Ferrari, D. Marioli, A. Taroni and M. Mazzoni, Comparisons between PZT and PVDF thick film technologies in the design of low cost pyroelectric sensors, Rev. Scientif. Instru. 75, 4906-4910 (2004). Cuadras, A., M. Casulla, A.Ghisla and V. Ferrari, Energy harvesting from PZT pyroelectric films, IMTC 2006 Instrumentation and Measurement Technology Conference, 1668-1672 (2006). Cuadras, a., M. Gasulla and V. Ferrari, Thermal energy harvesting through pyroelectricity, Sens. Actuators A 158, 132-139 (2010). Vanderpool, D., J. H. Yoon and L. Pilon, Simulations of a prototype device using pyroelectric materials for harvesting waste heat, Int. J. Heat Mass Transfer 51, 5052-5062 (2008). Navid, A., D. Vanderpool, A. Bah and L. Pilon, Towards optimization of a pyroelectric energy converter for harvesting waste heat, Int. J. Heat Mass Trans. 53, 4060-4070 (2010). Nguyen, H., Navid, A. and L. Pilon, Pyroelectric energy converter using co-polymer PVDF-TrFE and Olsen cycle for waste heat energy harvesting, Appl. Therm. Engin. 30, 2127-2137 (2010). Navid, A. and L. Pilon, Pyroelectric energy harvesting using Olsen cycles in purified and porous PVDFTrFE thin films, Smart Mater. Struct. 20, 025012 (2011). M. Ujihara, G. P. Carman and D. G. Lee, Thermal energy harvesting device using ferromagnetic materials, Appl. Phys. Lett. 91, 093508 (2007). J. Claude, Y. Lu, K. Li and Q. Wang, Electrical storage in poly(vinlyidene fluoride) based ferroelectric polymers, Chem. Mater. 20, 2078-2080 (2008). J. Li, S.I. Seok, B. Chu, F. Dogan, Q. Zhang abd Q. Wang, Nanocomposites of ferroelectric polymers with TiO2 nanoparticles exhibiting significantly enhanced electrical energy density, Adv. Mat . 21, 217221 (2009). J. Wie, P. P. Mane, C. W. Green, K. M. Mossi and K. K. Leang, Energy harvesting by pyroelectric effect using PZT, Proc. ASME Conf. Smart Materials, Adaptive Structures Intelligent Systems , Ellicott City, MY, 1-5 (2008). H. H. S. Chang and Z. Huang, Substantial pyroelectric effect enhancement in laminated composites, Appl. Phys. Lett. 92, 152903 (2008).

12

[40]

Lavrik, N.V., M.J. Sepaniak, P.G. Datskos, Cantilever transducers as a platform for chemical and biological sensors, Rev. Scientif. Instru. 75, 2229-2253 (2004).

13

You might also like