Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

*Revised Manuscript with No Changes Marked

Click here to download Revised Manuscript with No Changes Marked: Main_manuscrip_supplementary_doc_not_marked.docx

1
2
3
4
5 Shape memory alloy engine for high efficiency low-temperature gradient
6
7 thermal to electrical conversion
8
9 Prashant Kumara*, Ravi Anant Kishorea,e, Deepam Mauryaa, Colin J Stewarta, Reza Mirzaeifarb,
10
11 Eckhard Quandtc, and Shashank Priyaa,d*
12
13 a
14 Center for Energy Harvesting Materials and Systems (CEHMS), Virginia Tech, 310 Durham
15
16
Hall, Blacksburg, VA 24061, USA
17
b
18 Department of Mechanical Engineering, Virginia Tech, Blacksburg, VA 24061, USA
19
20 c
21 Kiel University, Institute for Materials Science, Kaiserstr. 2, 24143 Kiel, Germany
22
d
23 Materials Research Institute, Penn State, University Park, PA 16802.
24
25 e
26 National Renewable Energy Laboratory, 15013 Denver West Pkwy, Golden, CO 80401, USA.
27
28 Abstract
29
30
31 More than half of the energy generated worldwide is lost as unused thermal energy because of
32
33 the lack of efficient methodology for harnessing the low-grade heat. Here we demonstrate that
34 shape-memory alloy can be an effective mechanism for recovering low-grade heat. Shape
35
36 memory alloys exhibit thermally induced martensite to austenite phase transformation and super-
37
38 elasticity (stress-induced martensitic transformation). Employing these two characteristics, we
39
40 demonstrate a thermal engine for harnessing waste energy through all modes of heat transfer:
41
42 convection, conduction, and radiation. In this work, we performed material and heat transfer
43
44
analysis for achieving high frequency, sustainable and efficient operation of our engine. An
45 optimized shape memory alloy engine generated 36 W from 1 kg or 234 kW of electricity from 1
46
47 m3 of active material. A continuous three-day operation of several SMA engines could generate
48
49 7.2 kWh of electricity when installed on a 500 m long hot pipe network. This generated power
50
51 can reduce the carbon footprint by 5.1 kg of CO2 illustrating the promise of this technology for
52
53 addressing climate change.
54
55 Keywords: Shape Memory Alloy (SMA); Martensite; Austenite; Pseudo-elasticity; Heat engine;
56
57 Energy harvesting.
58
59
60 *Corresponding author: pkumar14@vt.edu, sup103@psu.edu
61
62 1
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 Introduction
5
6
7 Among the various forms of environmental energy available around us, thermal energy is the
8
9 most abundant and ubiquitous. Thus, there have been considerable efforts made towards
10
11 developing techniques for conversion of thermal energy into electrical energy. Thermal energy is
12 usually classified as high-grade, medium-grade and low-grade based in its temperature [1]. As
13
14 shown in Fig.1(a), low-grade waste heat provides a huge work potential as it is abundantly
15
16 present around us. In most of the industrial processes, a large amount of waste heat is generated,
17
18 where, a major portion of this thermal energy is available in the form of low grade waste heat
19
20 [2]. Harvesting this low grade thermal energy is highly desirable for efficient industrial process
21
22
and environmental impact [3]. The thermal to electrical conversion process, however, becomes
23 complex with the decrease in heat source temperature. The traditional steam Rankine cycle based
24
25 power plants are currently the most effective technology to obtain work from heat. However,
26
27 these systems are bulkier and not efficient for low grade heat energy harvesting [4-6]. Other
28
29 alternatives such as Organic Rankine cycle (ORC) and Kalina cycle have been deployed for low
30
31 temperature heat applications and waste heat recovery [7, 8]. The difficulty in obtaining suitable
32
33
organic fluid for ORC [9], and proprietary nature of Kalina cycle [10] limit their practical
34 exploitation, especially for small-scale applications. In some of the methods, heat upgrading
35
36 techniques (heat pumps and absorption heat transformers (AHTs)) are deployed to utilize the low
37
38 grade waste heat [2, 11]. However, reliability of these technologies and their environmental
39
40 effects are some of the challenges need to be solved [12].
41
42 In recent years, significant efforts have been made to explore and develop the material based
43
44 alternative technologies for low-grade thermal energy harvesting and waste heat recovery, such
45
46 as thermoelectric [13-15], pyroelectric [16-18], thermomagnetic [19, 20], thermo-acoustics [21],
47
48 and thermo-electrochemical [22, 23]. In thermoelectric generators (TEGs), a direct thermal to
49
50 electrical energy conversion takes place due to the Seebeck effect and the performance relies on
51
52
figure of merit of the material [24]. Pyroelectric devices are used to convert the fluctuating
53 temperature directly into electricity using ferroelectric materials [18]. Thermomagnetic device
54
55 uses magnetic materials which undergo secondary phase change on thermal loading. This
56
57 phenomenon is used to convert thermal to mechanical to electrical energy [23, 25]. In thermo-
58
59 acoustic based device, a temperature gradient is used to produce acoustic waves by utilizing
60
61
62 2
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 thermal energy, and acoustic wave is then converted to electricity by using piezoelectric material
5
6 [21]. In another example, low temperature waste heat is directly converted into electricity
7
8 through electrochemical cycle by utilizing electrode materials with low heat and high charge
9
10 capacity. Out of all these material based choices, TEGs have dominated the scientific interest in
11
12 capturing locally available thermal energy. However, there is a significant drop in the
13
14
performance of TEGs (efficiency ~1-3%), when hot-side temperature is below 100 oC [26].
15 Other material based techniques mentioned above in current form provide smaller output power
16
17 density and thus remain early-stage laboratory research. In trying to address this decades old
18
19 grand challenge, we made a breakthrough in demonstration of small-scale heat engine based on
20
21 shape memory alloy (SMA). SMA based engine was designed to operate at temperatures less
22
23 than 80oC with the ambient acting as heat sink. The engine relies on two fundamental properties
24
25
of SMA: (i) super-elasticity, and (ii) thermally induced martensite to austenite phase
26 transformation.
27
28
29 There have been several attempts to develop the SMA system for converting heat energy into
30
31 output mechanical work[27-29]. Recently, Sato et al. [30] have presented the large scale working
32
33
device based on SMA that demonstrated 1.155 W output power for 40.25 cm3 of active material
34 volume (5 belt, weight of active material ~ 0.262 kg, device dimension: 18.50 cm x 5.50 cm x 5
35
36 cm). However, most of these previous designs have remained laboratory experiments and their
37
38 reliability and durability for a long-term domestic or commercial application remains
39
40 challenging. Especially for the devices requiring rotation, challenges arise from the fact that heat
41
42 needs to be captured from the source (hot-side temperature less than 80oC) at a very fast rate
43
(several hundred rotations per minute) through an extremely thin interface (SMA wire diameter
44
45 less than few hundred microns). Also, the residual heat in the wire has to be completely
46
47 discarded to the ambient at equally fast rate to achieve continuous operation. As the size of SMA
48
49 engine is reduced, the hot-side and the cold-side come closer to each other, which results in
50
51 continuous accumulation of heat after each cycle. Eventually the accumulated heat stops the
52
53 functioning of the device due to insufficient cooling. This has been the challenge towards
54 realizing a small scale SMA engine for the past four decades (since 1975).
55
56
57 Here, we provide a breakthrough in developing small scale SMA engines operating at hot-
58
59 side temperatures less than 100 °C and overcoming all of the above mentioned challenges. The
60
61
62 3
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 potential of our SMA engine can be recognized from the data presented in Fig. 1(a). This data
5
6 shows the amount of waste heat available corresponding to different thermal gradients and low
7
8 grade heat (230 °C) is the major component of total waste heat. Fig. 1(b) shows the geothermal
9
10 locations (wells and springs) [31] across the United States, which could provide the opportunity
11
12 for deployment of SMA engine arrays demonstrated in this study. Fig. 1(c) shows various
13
14
locations of hot pipes in the industrial and residential settings with low thermal gradient waste
15 heat.
16
17
18 For harvesting this abundant amount of low grade thermal energy, the SMA wire needs to
19
20 possess a low transition temperature, smaller hysteresis, low heat capacity, high thermal
21
22
conductivity, and high thermo-elastic efficiency. Therefore, extensive investigations were
23 conducted on phase transition behavior, thermodynamic properties, and thermal hysteresis of
24
25 SMA wires. Using the measured SMA material characteristics, we designed an engine
26
27 comprising of two pulleys with different diameters, a thin SMA belt around two pulleys, a
28
29 metallic container for fluid storage, and a small DC electric generator (Fig. S1). Fig. 1(d) shows
30
31 array of SMA engines deployed on hot pipes (Fig. 1(c)). In this scenario, the heat from hot pipe
32
33
will be conducted to fluid through metallic base of the container. The hot fluid then heats the
34 SMA wire above the transition temperature to run the engine. Our SMA engine was found to
35
36 generate sufficient power needed for powering water health monitoring sensors and acoustic
37
38 devices with hot-side temperature ranging between 60-80 °C. It is worthwhile to mention that the
39
40 engine design presented here can be used at much lower hot-side temperatures (10 °C above the
41
42 ambient) if the transition temperature of SMA can be reduced by modifying its composition [32].
43
The long–term dynamic mechanical analysis (continuous three day operation) on the SMA wire
44
45 indicated no significant change in the thermo-mechanical properties. The calculation presented in
46
47 the supplementary file, projects that our SMA engine can generate 7.2 kWh over three-day
48
49 period in an industrial setup comprising of 500 m long hot pipe (at ~80-90oC) network. This
50
51 generated power can reduce the carbon footprint by 5.1 kg of CO2.
52
53
54 Methods and Experiments
55
56
The experiments consisted of three main steps. In the first step, differential scanning
57 calorimetry (DSC) and dynamic mechanical analyzer (DMA) studies were performed on five
58
59 different SMA wires and based upon the results most optimum wire was selected. Five samples
60
61
62 4
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 of NiTi were used in as-received form for the DSC test: Wire 1 (0.25 mm, Muscle wires), Wire 2
5
6 (0.38 mm, DYNALLOY, Inc), Wire 3 (0.2 mm, Johnson Matthey), Wire 4 (0.38 mm, DYNALLOY,
7
8 Inc), and Wire 5 (0.44 mm, Sci-supply)). The SMA wires were cut into small pieces and the test
9
10 specimens were washed thoroughly with acetone to remove any surface impurities. Before
11
12 recording the final readings, each test specimens were heated and cooled multiple times in the
13
14
temperature range of -40 °C to 100 °C in a closed furnace with heating/cooling rate of 10
15 °C/min. Further, DMA tests were conducted on the selected SMA wire. During these
16
17 measurements, the sample was scanned from -20 °C to 100 °C at different temperature scan rates
18
19 (1 °C/min, 2 °C/min and 3 °C/min) and frequencies (0.1 Hz,1.0 Hz, and 10 Hz). In the second
20
21 step, dynamic thermal analysis was performed on the selected wire. The goal of the experiment
22
23 was to identify the temperature distribution of the moving wire during operation, right from the
24
25
point where the wire moves out of the contact with the heat source to the point where it returns
26 back to make contact with the heat source. This experiment is extremely important to understand
27
28 the thermal response of the SMA wire vis-à-vis thermoelastic cycle efficiency of the SMA
29
30 material when it is pre-stressed between the pulleys. The detailed schematic of the experimental
31
32 setup is shown in Fig. S2(a). Since the SMA wire was thin (~ 0.44 mm) and moving, the
33
34 experiment was very sensitive to external factors such as ambient temperature and air speed.
35
Therefore, the entire experimental set-up was thermally isolated during the measurements.
36
37 Infrared (IR) camera (FLIR SC6700, FLIR Systems, Inc.) was used to capture the radiation
38
39 coming from the thin moving SMA wire and the temperature gradient was evaluated
40
41 accordingly. Before running the experiments, IR camera was calibrated with a thermocouple by
42
43 matching the surface temperature of the stationary wire under ambient condition (at 22°C). In
44
45 order to mitigate the experimental error due to the reflections from the surroundings, the entire
46 experimental setup was painted black and covered with black fabrics. In addition, in order to
47
48 obtain the accurate and high resolution thermal data, the entire engine was divided into small
49
50 sections and the IR radiation coming from each section was captured separately. Also, a thin
51
52 graphite layer was deposited on the SMA wire in order to enhance its surface emissivity. Before
53
54 capturing any data from the IR camera, a preliminary thermal analysis (DSC) was conducted to
55
56
understand the effect of the graphite layer on SMA wire. Fig. S2(b) shows the DSC results
57 before and after the graphite coating indicating very minor differences. Inset shows the scanning
58
59 electron microscope (SEM) image having a graphite coating. All the thermal videos representing
60
61
62 5
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 the dynamics of wire were processed using a MATLAB code to obtain the temperature profile.
5
6 In order to minimize the inconsistency during experiments, temperature data from several videos
7
8 were averaged and a final temperature gradient was obtained along the length of the wire. In
9
10 addition, a high resolution camera was used to capture the vibration of the moving wire along
11
12 with the thermal videos. The high resolution videos were then processed using MATLAB to
13
14
obtain vibrational characteristics such as amplitude and frequency.
15
16 In the third and last step, we measured the power output of SMA engine. Fig. S3 shows the
17
18 schematic diagram of the experimental set-up used to measure the mechanical and electrical
19
20 power output. Two pulleys of different diameters (optimized diameter ratio 1:3) were fixed on
21
22
two ends of 3-D slider stand through two small ball bearings. The SMA wire was looped around
23 the two pulleys. The optimal distance of 16.51 cm was maintained between the two pulleys. The
24
25 shaft through the lower pulley was connected to a small permanent magnet DC generator (rated
26
27 at 6V and Model # RF-500TB). A small portion of the SMA wire maintaining contact with the
28
29 lower pulley was heated either through hot water bath or a heat gun or radiation from a hot plate.
30
31 In order to determine the mechanical power, we first decoupled the electrical generator from
32
33
SMA engine. The angular speed of pulley was then recorded from the instant rotation starts to
34 the time rotation reaches the steady state. Moment of inertia (I) of the pulleys was obtained using
35
36 a CAD software, SOLIDWORKS.
37
38
39 The generator (coupled with engine) was connected to a resistance box (RS-201, IET LABS
40
41
Inc.) whose resistance could be varied between 0 to 500 Ω to identify the optimal load. A data
42 acquisition system, SIGLAB-SIGDEMO was used to acquire the voltage waveform. A voltmeter
43
44 (FLUKE 179) was also connected in parallel to the external resistance to monitor the output
45
46 voltage. An optical tachometer (SHIMPO DT-209X) was used to measure the angular speed of
47
48 the rotating pulleys.
49
50 Results and discussion
51
52
53 In the first stage, we performed material investigations to identify the most suitable SMA
54
55 alloy composition for engine design. Focus in this stage was on understanding the fundamental
56
57
material behavior under cyclic temperature and stress variations. In the second stage, we
58 performed thermal transport analysis to ensure rapid heat transfer rate across the SMA wire.
59
60
61
62 6
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 Lastly, in the third step, we performed systematic experiments on heat engine to fully quantify
5
6 the device performance.
7
8
9 SMA investigations for design of heat engine
10
11 A comprehensive DSC analysis was performed on different SMA wires to understand the
12
13 thermal deformation cycle and identify the composition which has maximum thermodynamic
14
15 efficiency at low temperatures under stress free condition. Supplementary Fig. S4(a)-(e) show
16
17 the hysteresis across phase change and the transition temperature for each sample. Using this
18
19 experimental heat flow diagram, the critical temperatures Ms (martensite start), Mf (martensite
20
finish), As (austenite start), and Af (austenite finish) for forward and reverse transformation
21
22 were quantified. It can be seen from Fig. S4 that wire 5 has the lowest forward transition
23
24 temperature of 48°C, which is most suitable for our heat engine. The difference between thermal
25
26 energy going-in and coming-out of the wires was found to be very small, which should be the
27
28 case since the DSC tests were run under no stress condition (no work). Residual heat
29
30 accumulation may cause thermo-mechanical fatigue in the system. Fig. S5 (a), (b) and (c) show
31
detailed results on phase transition of the annealed sample. Thermal annealing results in shift of
32
33 transition temperature from 48 °C to 54 °C. SMA wire was also characterized using dynamic
34
35 mechanical analyzer (DMA) in order to determine the force dynamics and viscoelasticity
36
37 properties. Detailed comparative study of damping and storage modulus values for as-received
38
39 and annealed SMA samples are shown in Supplementary Fig. S6-8. We observed that with the
40
41 increase of frequency the damping coefficient decreases, and with the increase of temperature
42 scan rate the transition temperature of the wire increases.
43
44
45 Thermal analysis of the heat engine
46
47
48
The power output of an SMA engine primarily depends on the thermoelastic cycle frequency
49 i.e. the rate at which the phase transition occurs in forward and reverse directions. The efficiency
50
51 of SMA engine can be enhanced by avoiding excessive heating or cooling during the thermal
52
53 cycle. In order to analyze the temperature distribution along the wire under operating conditions,
54
55 thermal videos were recorded for various sections of the wire, as shown in Fig. 2(a), while
56
57 maintaining the wire speed at 0.25 m/s. The first section was considered 1 cm above the hot
58 pulley to avoid the transient effects of thermal zone generated from hot source which assisted in
59
60
61
62 7
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 measuring the correct temperature values on wire surface. Detailed experimental configuration is
5
6 described in Fig. S2.
7
8
9 The vibration of SMA wire plays an important role in determining the performance of heat
10
11 engine, as it enhances the convective cooling of the wire. The convective heat transfer
12 coefficient, , in this case, depends on the velocity of the wire in the direction of rotation as well
13
14 as vibrational speed in the normal direction. Fig. S9(a) and (b) show the vibration profile
15
16 (captured by high resolution camera) of wire in the time domain and qualitative Fast Fourier
17
18 Transformation (FFT) response. The wire is vibrating without any fixed pattern and the
19
20 frequency response splits into many major frequencies between, ~ 0 to ~15 Hz.
21
22 Fig. 2(b) compares the temperature across a selected section of wire obtained experimentally
23
24 and empirically using equation (s7) for different values of heat transfer coefficient. It can be
25
26 noted that at h=250 W/m2-K, analytical results are in close agreement with the experimental
27
28 results (maximum deviation of less than 3%). Fig. S10 shows the heat transfer analysis of
29
30 moving wire and Supplementary Table S1 lists the thermal properties. This predicted heat
31 transfer coefficient includes the simultaneous effect of forced convection due to linear speed
32
33 (due to the rotation of the pulley) and vibration of the wire. Validated analytical model was used
34
35 to evaluate the temperature profile of the wire for heat transfer coefficients ranging from 200-400
36
37 W/m2-K, using Supplementary Equation (s7). As cooling coefficient approaches 320 W/m2-K
38
39 (with the wire speed of 1.8 – 2 m/s), the temperature of wire reduces by 3.75 °C more (compare
40
41
to h=200 W/m2-K) in the first section of the wire. This analysis indicates the location of wire
42 where the reverse phase transition temperature is achieved. We use this information to optimize
43
44 the length of the wire in order to reduce the size of heat engine.
45
46
47 Fig. 2(c) shows the temperature profile of wire loop starting from the emerging point near the
48
49 lower pulley to the upper pulley and from the upper pulley to lower pulley. Fig. 2(d) compares
50 the temperature profile obtained using our empirical model (Supplementary Equation (s7)),
51
52 experimental results, and the results obtained from the model proposed by Kase et al.[33]. Our
53
54 results differ from the model prediction of Kase et al. by 6-7%, which can be correlated with the
55
56 difference in the experimental setup and vibrational dynamics of the wire that includes the wire
57
58 curvature effect. Fig. 2(d) shows the effect of curvature near the upper cold pulley on the
59
60
temperature profile of wire.
61
62 8
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 SMA heat engine performance
5
6
7 Next, we quantified the performance of the engine by cycling the heat source (water)
8
9 temperature from 55-85 °C. The martensite to austenite transition completes at 54 °C, thus, the
10
11 hot-side temperature was maintained above 55 °C. The heat sink temperature was fixed at
12 ambient temperature. As shown in Fig. 3(a), the angular velocity of pulley varies in proportion
13
14 with hot side temperature. It is important to note that in Fig. S11, the angular speed of pulley
15
16 does not follow the same path during the heating and cooling cycle. Another important parameter
17
18 that affects the system’s efficiency was found to be the dip angle, the angle made at the center by
19
20 portion of the lower pulley exposed to heat source. Fig. 3(b) compares the variation in angular
21
22
speed with time at three different dip angles when the heat source is fixed at 70°C. It is
23 interesting to note that there exists an optimal dip angle (~ 56o) for maximum angular speed. The
24
25 angular speed was found to first increase with an increase in dip angle due to the increased heat
26
27 inflow. The angular speed is maximum when the dip angle is approximately equal to contact
28
29 angle of SMA wire with the pulley. Increasing the dip angle further reduces the angular speed.
30
31 This could happen because of increase in drag forces with increase in dip angle.
32
33 An optimized SMA engine with respect to heat source temperature and dip angle was
34
35 examined under different heating modes: hot water bath, hot air, and radiative heating from a hot
36
37 plate. Fig. 3(c) and (d) show the SMA engines speed (rotation per minute) operating under
38
39 different heat sources. The temperature of hot air was maintained at 125 °C. The hot plate
40
41
temperature was fixed at 250oC and the gap distance between hot plate and wire was 1 cm. In
42 order to enhance the heat transfer, silver paste was applied on inside the groove of lower pulley.
43
44 It can be noted from Fig. 3(c), that the silver paste improves the angular speed of SMA engine.
45
46 The maximum angular speed of the lower pulley in the steady state was measured to be 164 rpm
47
48 and 268 rpm for hot air and hot plate, respectively. Fig. 3(d) shows the angular speed of the
49
50 upper pulley at different hot-side temperatures after the system has reached steady state. It can be
51
52
seen that the SMA engine achieves a maximum angular speed when hot-side temperature is 80
53 °C and the corresponding angular speed is 420 rpm. It should be noted that the speed shown on
54
55 y-axis of Fig. 3(c) is the rpm of the lower pulley; whereas, the rpm shown in Fig. 3(d) is the rpm
56
57 of the upper pulley, which is three times lower than that of the lower pulley. From these results it
58
59 can be concluded that the SMA engine performs best when the heat source is hot water. This is
60
61
62 9
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 expected as the heat transfer coefficient in case of water is much higher than that of air.
5
6 Therefore, in the remainder of this study, hot water is used as the heat source.
7
8
9 In rotational dynamics, mechanical power is given as the product of torque and angular
10
11 speed, and it can be expressed as [34]:
12
13 d
14 Pmech  I   (1)
dt
15
16
17 where I is the moment of inertia of pulleys, d/dt is angular acceleration of pulleys, and  is
18
19 angular speed (rad/s). Next, we measured variation of angular speed as a function of time for
20
21 upper and lower pulleys, as shown in Fig. 4(a). A sixth order polynomial function was used to
22
23
obtain a functional relationship that describes the angular speed of the pulley with respect to
24 time. The angular acceleration of the upper and lower pulleys can be determined by taking the
25
26 time derivative of the polynomial expression. Angular acceleration multiplied with the total
27
28 moment of inertia of the rotating body provides the torque. This torque was then used to
29
30 determine the mechanical power by using equation (1).
31
32 Fig. 4(b) and (c) show the angular speed, torque, and mechanical power of the upper and
33
34 lower pulleys, respectively, at hot water temperature ( ) of 69 °C. These figures depict that the
35
36 torque is maximum when the pulley is just about to rotate (start-up torque), it decreases as the
37
38 angular speed increases, and it diminishes to zero after the steady state is achieved. The
39
40 mechanical power is zero at the beginning (since the system is at rest), it increases with the
41
42 increase in angular speed until it gains a maximum value (when the product of torque and
43 angular speed is highest), and then it slowly decreases. Fig 4(d) depicts the total mechanical
44
45 power of the SMA engine at °C. The maximum mechanical power can be found to be
46
47 12.5 mW. Fig. 5(a) shows the comparison of mechanical power at two different hot water
48
49 temperatures: °C and 80 °C. The maximum mechanical power obtained by the engine
50
51 is 26mW at 80 °C, which is more than twice the mechanical power obtained at
52
53 °C. The specific mechanical power, calculated by dividing the mechanical power over the
54
55
mass of the active component (SMA wire in this case), was found to be 52 W/kg at °C.
56
57 In order to quantify the electrical power output, we connected a small permanent magnet DC
58
59 motor (rated at 6 V) with the shaft of the lower pulley. Detailed experimental setup is shown in
60
61
62 10
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 Fig. S3. Fig. 5(b) shows the variation of generator’s shaft rpm and external load resistance at
5
6 different heat source temperatures (hot water). It can be seen that at a fixed load the shaft rpm
7
8 increases with increase in the hot water temperature. At a fixed temperature, the rpm first
9
10 increases with increase in the load resistance and then saturates. Fig. 5(c) shows the output DC
11
12 voltage (V) and the output current (I) obtained for different load resistances at a fixed hot water
13
14
temperature of 80°C. V-I plots at other hot water temperatures are not shown for the purpose of
15 graphical clarity. The voltage follows a similar trend as the rpm of the generator’s shaft, and the
16
17 maximum output voltage of 1.7 V was obtained at 80oC when the generator has virtually no load
18
19 (at 500Ω). Fig. 5(d) depicts the electrical power output as a function of external resistance at
20
21 different values of heat source temperature. It can be seen that SMA engine generates maximum
22
23 electrical power output of 18 mW across 70 Ω load resistance when hot water temperature is 80
24
25
°C. The electrical power can be scaled up by using multiple uncoupled devices connected to a
26 common heat source. To illustrate the scalability of our engine, we designed experiments with
27
28 three harvesting units (Fig. S12(a)) and the corresponding results are shown in Fig. S12(b). The
29
30 maximum output power can be obtained to ~24 mW at hot-side of 69oC.
31
32
33
Fig. 6 compares the efficiencies calculated for SMA heat engine. SMA material’s
34 thermodynamic efficiency is calculated to be 5.0% using Supplementary Equation (s8). The
35
36 absolute efficiency of SMA engine is 1.5%, which is 10.5% of the Carnot efficiency. It should be
37
38 noted that although output power increases with increase in hot water temperature, the efficiency
39
40 is maximum at 70 °C. This is an important observation because the forward phase transformation
41
42 occurs at 54 °C, which is effectively achieved when hot water temperature is 70 °C. Overheating
43
the SMA wire beyond this temperature decreases the system’s efficiency. The optimal heat
44
45 source temperature is the one where 100% forward phase transformation is completed. This is
46
47 also true for the heat sink temperature. Over-heating or under-cooling decrease the system’s
48
49 efficiency.
50
51
52
The effectiveness of the SMA heat engine was compared with other existing technologies
53 used for thermal-to-electrical energy conversion. In this domain, a thermoelectric generator
54
55 (TEG) is the most popular device. Normally, the output power and efficiency of TEGs are not
56
57 readily available for operation below 100oC. We, therefore, developed a numerical model using
58
59 ANSYS workbench v17.0 to determine these quantities and validated our calculations with the
60
61
62 11
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 experimental results reported by Hao et al. [13]. The numerical model was then used to produce
5
6 the power and efficiency data for TEG operating below 100 °C. Fig. S14 compares the numerical
7
8 results with the experimental results, which were found to be in close agreement. The numerical
9
10 model was then modified to account for ambient as the heat sink. From this analysis, at a
11
12 temperature above 70 °C, SMA engine is competitive (Fig. 7(a)) with TEG. Fig. 7(b) compares
13
14
different power indices in terms of active material power density and active material specific
15 power for TEG and SMA engine. The SMA engine can be seen to be competitive with TEG in
16
17 all the attributes. Further, we compared the cost of our engine with one of the commercially
18
19 available TEG and found that SMA engine is a much cheaper solution (supplementary, section
20
21 S10). Recently, thermal energy harvesting using tensile muscles has been reported [35], which
22
23 were shown to exhibit 7.2 W/kg at temperature difference of 70 oC. Comparatively, SMA engine
24
25
provides the specific power of 36 W/kg at lower temperature difference of 58 °C. In another
26 recent study, a small scale thermo-magnetic harvester has been realized for harvesting thermal
27
28 gradient by using phase transformations in gadolinium [19]. On the basis of information
29
30 available, we evaluated the specific realized power as ~0.6 W/kg for the temperature gradient
31
32 (ΔT) of 80 °C. This implies that the SMA engine presented in this study exhibits improved
33
34 performance compared to all current thermal energy harvesting technologies.
35
36 Demonstrations of SMA heat engine and future impact
37
38
39 We successfully demonstrated several practical applications which could be powered by
40
41
SMA heat engine. In the first demonstration, we continuously operated a sound projector in both
42 air and water medium (supplementary movie). Fig. 7(c) shows the sound generation in water
43
44 medium. The sound pressure is recorded by hydrophone (Brüel & Kjær type-8103). Fig. 7(d)
45
46 shows the waveform for generated acoustic pressure. This demonstration paves the pathway for
47
48 self-sustained air/under-water acoustic communication. We also demonstrate the real-time water-
49
50 health monitoring using the SMA heat engine (Fig. S15). This application provides direction for
51
52
the development of self-sustained health monitoring devices that can be deployed in residential
53 buildings and in natural environments such as hot springs.
54
55
56 Next, we evaluated the impact of our heat engine at large scale for a long duration of time.
57
58 Detailed calculations are shown in supplementary (section S13). Through DMA experiments
59
60
(Fig. S16), we show that the SMA can undergo thermal cycling for three days without any
61
62 12
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 degradation. On the basis of these outcomes, we predict that if our heat engine is operated for 1-3
5
6 days it can generate 2.4-7.2 kWh of energy (equivalent to 1.7 – 5.1 kg of CO2). To further
7
8 strengthen our arguments for long term operation, we operated heat engine for continuous 10
9
10 hours at 65oC in laboratory environment (Fig. S17), and did not observe any change in the
11
12 performance.
13
14 The durability of our system is related to the fact that phase transformation happens
15
16 instantaneously at each point of the material which experiences the required temperature
17
18 changes. However, in a bulk material, the temperature will not rise and drop uniformly over the
19
whole volume (i.e. the cross-section of wire). If the temperature changes are concentrated at the
20
21 surface of the wire, the whole material would not be contributing to the phase transformation,
22
23 and thereby, reduce efficiency. Also, a non-uniform distribution of phase transformation triggers
24
25 the fatigue failure. The observed efficiency and the fatigue resistance of the system further attest
26
27 the robustness of our engine. The selection of smaller radius wire and enhanced heating time,
28
29 ensured a uniform phase transformation (forward and reverse) in the wire (see supplementary,
30 section S14, for more discussion).
31
32
33 Conclusion
34
35 In summary, we demonstrate the operation of a low grade shape memory alloy heat engine
36
37 that operates below 80oC using ambient as the heat sink. Among the different heat sources
38
39 examined, the SMA engine performed best with the hot water as the heat source. The maximum
40
41 mechanical power of the shape memory alloy heat engine was found to be 12.5 mW at 69 °C and
42
43
26 mW at 80 °C. The mechanical power density of the engine was calculated to be 52 W/kg of
44 the active mass (shape memory alloy wire). The maximum electrical power of the SMA heat
45
46 engine was found to be 8.8 mW at 70 °C and 18 mW at 80 °C. The electrical power density of
47
48 the engine was calculated to be 36 W/kg of the active mass (shape memory alloy wire). The
49
50 shape memory alloy material’s thermodynamic efficiency was found to be 5.0%. The maximum
51
52 thermal-to-electrical conversion efficiency of the engine was 1.5%, which is 10.5% of the Carnot
53
54
efficiency. The engine was successfully demonstrated for powering acoustic projector and water
55 health monitoring sensors. Results show that shape memory alloy heat engines of the dimensions
56
57 shown here can reduce 1.7 kg of CO2 (carbon footprint) production per day.
58
59
60 Conflicts of interest
61
62 13
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 There are no conflicts to declare.
5
6 Acknowledgements
7
8
9 The authors would like to thank center for energy harvesting materials and systems (CEHMS)
10
11 for providing the facilities and infrastructure. S.P. acknowledges the financial support from
12 DARPA MATRIX program. P.K. acknowledges the funding from Office of Naval Research
13
14 (ONR). R.K. is supported through ICTAS Doctoral Scholarship and AMRDEC SBIR program.
15
16 D.M. acknowledges the support from Office of Basic Energy Science, Department of Energy.
17
18 We thank Dr. Bruce Orler for helping us in acquiring data from DSC and DMA.
19
20
21
22
23
24 Main figures:
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 14
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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 Fig. 1: Thermal energy available at various locations (industry, home, and geothermal sites) for
40 potential deployment of SMA based heat engine. (a) A bar chart representing the potential of
41
42 thermal energy harvesting in various categories (subdivided on the basis of temperature range)
43
44 [6]. The graph is redrawn on the basis of data given in reference 6. (b) Geothermal locations
45
46 across the United States (in temperature range of 55-100 oC) for the deployment of SMA
47
48 engine[31]. This image was created off the NREL geothermal prospector site. The underlying
49
50
data is compiled by the National Renewable Energy Laboratory for the U.S. Department of
51 Energy. (c) Various home and industrial hot pipe locations for effective utilization of SMA
52
53 engine. (d) An array of SMA engines deployed on pipe (along with single device) for the
54
55 prospective utilization of hot thermal zone for maximizing the power and reducing the carbon
56
57 foot print.
58
59
60
61
62 15
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42 Fig. 2: Dynamic thermal investigation of thin SMA wire (active material in the engine) by IR
43
44 thermography. (a) Basic design of the SMA engine. The active material is divided into many
45
46 sections for accurate thermal profiling of thin and vibrating wires. (b) Experimentally obtained
47
48
temperature profile of moving wire (in steady state) for 1st section of active material at particular
49 dynamic condition and its comparison with modeled temperature profile at different predicted
50
51 heat transfer coefficients (h). Temperature profile obtained at h=250 W/K-m2 is matching with
52
53 experiments. (c) Temperature profile for the all the sections where wire is cooled continuously
54
55 after emerging from hot source. Section just before hot source acts as transition zone for heating
56
57 of cooled wire. (d) Complete thermal cycle of engine from heating-cooling-heating, and
58 comparison of cooling profile with predicted h value and theoretical relationship developed by
59
60 S.Kase et al.[33].
61
62 16
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44 Fig. 3: Performance characterization of optimized SMA engine for energy harvesting purposes.
45
46 (a) SMA engine performance under load for increasing and decreasing the hot side temperature
47
48
continuously. This shows the increasing and decreasing trend of RPM with respect to
49 temperature. (b) Performance optimization of SMA engine at different dip angles (length of wire
50
51 exposed to hot bath). (c) Engine performance (in terms of lower pulley rotation) with different
52
53 forms of non-contact heat sources such as hot air and hot plates. (d) SMA engine performance
54
55 (in terms of upper pulley rotation) at different temperatures of hot bath. A significant increase
56
57 has been observed with increasing hot side temperature.
58
59
60
61
62 17
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37 Fig. 4: Mechanical power calculation for SMA engine. (a) Angular velocity of upper pulley
38
39 along with 6th order polynomial curve fit when hot water is maintained at 69°C. The angular
40
41 acceleration can be determined by taking the time derivative of the polynomial expression. (b)
42 Angular velocity, derived torque and mechanical power (considering the upper pulley) in time
43
44 domain. (c) Angular velocity, derived torque and mechanical power (considering the lower
45
46 pulley) in time domain. (d) Final evaluated mechanical power of whole SMA engine when hot
47
48 water bath is at 69°C. The maximum mechanical power is around 12.5 mW.
49
50
51
52
53
54
55
56
57
58
59
60
61
62 18
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44 Fig. 5: SMA engine performance. (a) Mechanical power for two different thermal input
45
46 conditions. Maximum mechanical power obtained is 52W/kg (26mW) for operating range
47
48 considered in this study. (b) Generator (motor operated in reverse) shaft rotation per minute
49
50
(rpm) at different load with different thermal input temperature from hot water bath. (c) Output
51 current and output voltage as a function of external load resistance for particular thermal input
52
53 condition. Maximum voltage obtained is ~1.7V (at 500 Ohms) for temperature considered here.
54
55 (d) Power obtained as a function of externally applied resistance at different thermal input.
56
57 Maximum electrical power obtained is 36W/kg (18mW) within the operating temperature range
58
59
60
61
62 19
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 at 69% of generator efficiency. The matching external resistance for maximum power output is
5
6 70Ω.
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46 Fig. 6: Comparison of different type of efficiencies for the SMA engine. Carnot efficiency is
47
48 calculated from heat source and sink temperatures. Absolute efficiency is system’s actual
49
50
efficiency and evaluated as electrical power output/ thermal energy input. Material
51 thermodynamic efficiency is deduced from the phase transition curve of SMA wire by using
52
53 fundamentals of thermodynamics. This gives the maximum material capacity for power
54
55 scavenging. Relative efficiency is defined as absolute efficiency/Carnot efficiency. SMA engine
56
57 performs maximum efficiency when hot water bath is at 70°C.
58
59
60
61
62 20
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45 Fig. 7: Comparison of SMA engine with existing state of the art thermal energy harvesting
46
47 technology such as TEG, and experimental setup for powering acoustic projector. (a) Simulation
48 results on TEG under different boundary conditions and comparison with the efficiency of basic
49
50 version of SMA engine. At temperature above 70 °C, SMA engine is competitive with TEG. (b)
51
52 Various power density indices such as power per unit volume and power per unit mass of active
53
54 materials along with efficiency are compared for TEG and SMA engine. (c) Experimental setup
55
56 for demonstration of practical thermal energy harvesting for real-time under water acoustic
57
58
measurement. (d) Time domain acoustic pressure waveform generated in water from a small
59 SMA engine. The pressure was recorded at 2 cm above the device by hydrophone.
60
61
62 21
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
References:
11
12 [1] Kishore RA, Priya S. Low-grade waste heat recovery using the reverse magnetocaloric effect.
13 Sustainable Energy & Fuels. 2017;1:1899-908.
14 [2] Oluleye G, Jobson M, Smith R, Perry SJ. Evaluating the potential of process sites for waste
15
16 heat recovery. Applied Energy. 2016;161:627-46.
17 [3] Brückner S, Liu S, Miró L, Radspieler M, Cabeza LF, Lävemann E. Industrial waste heat
18 recovery technologies: an economic analysis of heat transformation technologies. Applied
19 Energy. 2015;151:157-67.
20
21
[4] Hettiarachchi HM, Golubovic M, Worek WM, Ikegami Y. Optimum design criteria for an
22 organic Rankine cycle using low-temperature geothermal heat sources. Energy. 2007;32:1698-
23 706.
24 [5] Barbier E. Geothermal energy technology and current status: an overview. Renewable and
25 Sustainable Energy Reviews. 2002;6:3-65.
26
27 [6] Johnson I, Choate WT, Davidson A. Waste Heat Recovery. Technology and Opportunities in
28 US Industry. BCS, Inc., Laurel, MD (United States); 2008.
29 [7] Bao J, Zhao L. A review of working fluid and expander selections for organic Rankine cycle.
30 Renewable and sustainable energy reviews. 2013;24:325-42.
31
[8] Shi L, Shu G, Tian H, Deng S. A review of modified Organic Rankine cycles (ORCs) for
32
33 internal combustion engine waste heat recovery (ICE-WHR). Renewable and Sustainable Energy
34 Reviews. 2018;92:95-110.
35 [9] Tocci L, Pal T, Pesmazoglou I, Franchetti B. Small scale organic rankine cycle (ORC): A
36 techno-economic review. Energies. 2017;10:413.
37
38
[10] Bombarda P, Invernizzi CM, Pietra C. Heat recovery from Diesel engines: A
39 thermodynamic comparison between Kalina and ORC cycles. Applied Thermal Engineering.
40 2010;30:212-9.
41 [11] Oluleye G, Smith R, Jobson M. Modelling and screening heat pump options for the
42 exploitation of low grade waste heat in process sites. Applied energy. 2016;169:267-86.
43
44 [12] Chua KJ, Chou SK, Yang W. Advances in heat pump systems: A review. Applied energy.
45 2010;87:3611-24.
46 [13] Hao F, Qiu P, Tang Y, Bai S, Xing T, Chu H-S, Zhang Q, Lu P, Zhang T, Ren D, Chen J.
47 High efficiency Bi 2 Te 3-based materials and devices for thermoelectric power generation
48
49
between 100 and 300° C. Energy & Environmental Science. 2016;9:3120-7.
50 [14] Zheng X, Liu C, Yan Y, Wang Q. A review of thermoelectrics research–Recent
51 developments and potentials for sustainable and renewable energy applications. Renewable and
52 Sustainable Energy Reviews. 2014;32:486-503.
53 [15] Elsheikh MH, Shnawah DA, Sabri MFM, Said SBM, Hassan MH, Bashir MBA, Mohamad
54
55 M. A review on thermoelectric renewable energy: Principle parameters that affect their
56 performance. Renewable and Sustainable Energy Reviews. 2014;30:337-55.
57 [16] Ravindran S, Huesgen T, Kroener M, Woias P. A self-sustaining micro thermomechanic-
58 pyroelectric generator. Applied Physics Letters. 2011;99:104102.
59
60
61
62 22
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 [17] Sebald G, Guyomar D, Agbossou A. On thermoelectric and pyroelectric energy harvesting.
5
6 Smart Materials and Structures. 2009;18:125006.
7 [18] Sebald G, Lefeuvre E, Guyomar D. Pyroelectric energy conversion: optimization principles.
8 IEEE transactions on ultrasonics, ferroelectrics, and frequency control. 2008;55.
9 [19] Chun J, Song H-C, Kang M-G, Kang HB, Kishore RA, Priya S. Thermo-Magneto-Electric
10
11
Generator Arrays for Active Heat Recovery System. Scientific Reports. 2017;7:41383.
12 [20] Elliott J. Thermomagnetic generator. Journal of Applied Physics. 1959;30:1774-7.
13 [21] Smoker J, Nouh M, Aldraihem O, Baz A. Energy harvesting from a standing wave
14 thermoacoustic-piezoelectric resonator. Journal of Applied Physics. 2012;111:104901.
15 [22] Abraham TJ, MacFarlane DR, Pringle JM. High Seebeck coefficient redox ionic liquid
16
17 electrolytes for thermal energy harvesting. Energy & Environmental Science. 2013;6:2639-45.
18 [23] Lee SW, Yang Y, Lee H-W, Ghasemi H, Kraemer D, Chen G, Cui Y. An electrochemical
19 system for efficiently harvesting low-grade heat energy. Nature communications. 2014;5:3942.
20 [24] Lee H, Sharp J, Stokes D, Pearson M, Priya S. Modeling and analysis of the effect of
21 thermal losses on thermoelectric generator performance using effective properties. Applied
22
23 Energy. 2018;211:987-96.
24 [25] Chun J, Kishore RA, Kumar P, Kang M-G, Kang HB, Sanghadasa M, Priya S. Self-Powered
25 Temperature-Mapping Sensors Based on Thermo-Magneto-Electric Generator. ACS applied
26 materials & interfaces. 2018;10:10796-803.
27
28
[26] Ismail BI, Ahmed WH. Thermoelectric power generation using waste-heat energy as an
29 alternative green technology. Recent Patents on Electrical & Electronic Engineering (Formerly
30 Recent Patents on Electrical Engineering). 2009;2:27-39.
31 [27] Wakjira JF. The VT1 shape memory alloy heat engine design: Virginia Polytechnic Institute
32 and State University; 2001.
33
34 [28] Schiller EH. Heat engine driven by shape memory alloys: prototyping and design: Virginia
35 Tech; 2002.
36 [29] Avirovik D, Kumar A, Bodnar RJ, Priya S. Remote light energy harvesting and actuation
37 using shape memory alloy—piezoelectric hybrid transducer. Smart Materials and Structures.
38
39
2013;22:052001.
40 [30] Sato Y, Yoshida N, Tanabe Y, Fujita H, Ooiwa N. Characteristics of a new power
41 generation system with application of a shape memory alloy engine. Electrical Engineering in
42 Japan. 2008;165:8-15.
43 [31] https://maps.nrel.gov/geothermal-prospector/.
44
45 [32] Villanueva A, Gupta S, Priya S. Lowering the power consumption of Ni-Ti shape memory
46 alloy. SPIE Smart Structures and Materials+ Nondestructive Evaluation and Health Monitoring:
47 International Society for Optics and Photonics; 2012. p. 83421I-I-12.
48 [33] Kase S, Matsuo T. Studies on melt spinning. I. Fundamental equations on the dynamics of
49
melt spinning. Journal of Polymer Science Part A: General Papers. 1965;3:2541-54.
50
51 [34] Kishore RA, Coudron T, Priya S. Small-scale wind energy portable turbine (SWEPT).
52 Journal of Wind Engineering and Industrial Aerodynamics. 2013;116:21-31.
53 [35] Kim SH, Lima MD, Kozlov ME, Haines CS, Spinks GM, Aziz S, Choi C, Sim HJ, Wang X,
54 Lu H, Qian D, Madden JDW, Baughman RH, Kim SJ. Harvesting temperature fluctuations as
55
56
electrical energy using torsional and tensile polymer muscles. Energy & Environmental Science.
57 2015;8:3336-44.
58
59
60
61
62 23
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12 Shape memory alloy engine for high efficiency low-temperature gradient
13
14 thermal to electrical conversion
15
16
17 Prashant Kumara*, Ravi Anant Kishorea,e, Deepam Mauryaa, Colin J Stewarta, Reza Mirzaeifarb,
18
19 Eckhard Quandtc, and Shashank Priyaa,d*
20
21 a
Center for Energy Harvesting Materials and Systems (CEHMS), Virginia Tech, 310 Durham
22
23 Hall, Blacksburg, VA 24061, USA
24
25 b
26 Department of Mechanical Engineering, Virginia Tech, Blacksburg, VA 24061, USA
27
28 c
Kiel University, Institute for Materials Science, Kaiserstr. 2, 24143 Kiel, Germany
29
30 d
31 Materials Research Institute, Penn State, University Park, PA 16802.
32
33 e
National Renewable Energy Laboratory, 15013 Denver West Pkwy, Golden, CO 80401, USA.
34
35
36
37
38
39
40
41
42
Supplementary Information
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 24
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
S1. Basic schematics of the system and working principle:
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. S1: (a) Schematic of front and side views of shape memory alloy engine. SMA wire gets the
35 heat from lower pulley (dipped in hot bath). Due to phase transition of SMA wire on thermal
36
37 input, strain develops inside the wire (contracts), and produces the torques in pulley (due to
38 contact friction between wire and pulley). This torque then turns the pulley and attached
39 generator and produces the electricity. (b) Schematic representation of displacement vs.
40
41 temperature profile of shape memory alloy. With increase in temperature, displacement
42 decreases and generates the differential tension across the pulley
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 25
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
S2. Thermal analysis setup
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Fig. S2: (a) Schematic showing the setup for thermal analysis of thin dynamic shape memory
54 wire (graphite coated) wrapped around pulleys. IR thermography has been used to capture the
55 temperature. (b) Differential scanning calorimetry (DSC) results for SMA wire, with and without
56
57 graphite coating. The differences are small, therefore, any significant effect on wire behavior
58 during thermal analysis has been neglected.
59
60
61
62 26
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
S3. Experimental setup for converting thermal to electrical energy:
8
9
10
11
12
13
14
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
40
41
42
43 Fig. S3: Schematic showing the experimental setup for quantifying the conversion of thermal
44
45 energy into electrical energy. Electrical power at different temperatures were measured.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 27
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S4. Differential scanning calorimetry (DSC) results for as-received wires:
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Fig. S4: DSC results for five different wires from various companies. (a) Wire 1 (0.25 mm,
54 Muscle wires). (b) Wire 2 (0.38 mm, DYNALLOY, Inc). (c) Wire 3 (0.2 mm, Johnson Matthey).
55
56 (d) Wire 4 (0.38 mm, DYNALLOY,Inc). (e) Wire 5 (Sci-supply). These results were important
57 to find the suitable wire.
58
59
60
61
62 28
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S5. Differential scanning calorimetry (DSC) results for annealed wires:
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53
Fig. S5: DSC results for annealed SMA wire (selected for SMA engine) at different
54 temperature scanning rate; 5,10 and 20°C/min. Annealing was done at 600 oC for 30 minutes,
55 (a) complete cycle of heating and cooling of sample, (b) and (c) enlarged graphs of heating
56 and cooling profile, and the comparison at different temperature ramping rate.
57
58
59
60
61
62 29
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S6. DMA analysis for determining the dynamic thermo-mechanical properties of wires:
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
Fig. S6: DMA results for un-annealed selected NiTi wire. (a) Tan delta measurement of SMA
60 wire for three frequencies. (b) Tan delta measurement of SMA wire for three temperature rate.
61 (c),(d),(e) and (f) shows the comparison of tan delta results of wire for different temperature
62 30
scan rates and frequencies.
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 Fig. S7: DMA results for un-annealed selected NiTi wire, (a),(b),(c),(d),(e)and (f) shows the
55 comparison of storage modulus results for different temperature scan rates and frequencies.
56
57
58
59
60
61
62 31
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
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
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 Fig. S8: DMA results for annealed selected NiTi wire, (a),(b),(c),(d),(e)and (f) shows the
56 comparison of storage modulus results for different temperature scan rates and frequencies.
57
58
59
60
61
62 32
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4
5
6
S7. Thermal modeling of rotary SMA engine:
7
8
9
10
11 0.006
12 (a) Vibration of Wire
(b)
13 Mean Position
0.004
Amplitude of Vibration (m)

14
15

Amplitude (a.u)
16
0.002
17
18
19
0.000
20
21
22 -0.002
23
24
25 -0.004
26 0 2 4 6 8 10 12 0 2 4 6 8 10 12 14
27
Time in Seconds (s) Frequency (Hz)
28
29
30
Fig. S9: (a) Horizontal vibration profile of wire during system operations. (b) The time domain
31 vibration profile is converted into frequency domain to observe the dominant frequencies.
32
33
34
35
36
37
38
39
x0 x0
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Fig. S10: Conceptual diagram for modeling of moving wire. For simplicity of modeling, wrapped
54 wire is converted into linear moving wire.
55
56
57
58
59
60
61
62 33
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 The exact analytical thermal modeling of rotary SMA engine is complex due to curvature effect
5
6 in heat transfer, dynamic metal to metal contact heat transfer, non-linear response to thermal
7 input etc. However, a simplified solution can be obtained which provides realistic estimation for
8
9 SMA engine behavior. For reducing the modeling complexity, we converted the lopped wire into
10 linear profile and allowed it to move through hot source as shown in Figure S10. In our model, a
11 wire is moving from negative direction to positive direction and passing through heat source
12
13 placed at origin. Heat transfer analysis on an element ( dx ) of moving wire is shown in equation
14 (s1):
15
16 dq
17 q q dx  mC p dT  dqc  dqr (s1)
18 dx
19
20 where q is heat flow, m is weight of wire, Cp is heat capacity of wire
21
22 Moving mass can be further given by equation (s2)
23
24 m   AV (s2)
25
26 where ρ is SMA density, A is cross section area of wire, V is velocity of wire.
27
28
29
Conductive heat transfers while heat source is at origin can be given by equation (s3)
30
dT
31
32
q   kA (s3)
33
dx
34
35 where k is thermal conductivity of SMA wire. The convective heat loss through moving element
36 of wire is given by equation (s4)
37
38 dqc  h(T  T )dA , and dA  Pdx (s4)
39
40
41 where h is heat transfer coefficient and T is ambient temperature. Thermal radiation component
42
43
is given by equation (s5) as:
44
45 dqr   (T 4  T 4  )dA (s5)
46
47 where  is Stefan–Boltzmann constant,  is emissivity. Since the temperature difference is not
48
49 that significant, the thermal radiation impact can be safely being ignored.
50
51 The heat equation for moving wire through stationary heat source is shown in equation (s6)
52
53 d 2T C pV dT hP  P 4 4
54 2
+  (T  T )  (T  T  )  0 (s6)
55 dx k dx kA kA
56
57 On solving the heat equation, an expression of distance dependent temperature is obtained as
58
shown in equation (s7)
59
60
61
62 34
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 
1  C pV  4 hkP  AC p 2V 2  2  
1  C pV  4 hkP  AC p 2V 2  2 
5  x   x 
6 2  k Ak  2  k Ak 
7 T ( x)  T  C1e  
 C2e  
(s7)
8
9 Constants C1 and C2 are determined using two boundary conditions, such as the temperature of
10
11 the wire at two locations (obtained experimentally from IR Thermography).
12
13 Figure S9 (a) and (b) show the vibration characteristic of the wire for mentioned thermal
14 experiment. As can be seen from the figures, wire has sufficient vibration in horizontal direction
15
16
which has significant contribution in overall cooling of the active material.
17
18 Thermal analysis parameters [2]:
19
20
21
22 Density Heat Capacity Thermal Conductivity Ambient Temperature
23
24
(Kg/m3) (J/(kg·K)) (W/m-K) (oC)
25 6450 322 18 22
26
27 Table S1
28
29
30
31 Table S1: Various important parameters which are used for modeling of thin moving SMA wire.
32
33 S8. System reversibility with respect to the temperature:
34
35
36 300
37
38
39 250 Cooling
Rotation per minute (RPM)

40
41
42 200
43
44
45
150
46
47
Heating
48
49 100
50
51
52 50
53
54
55
50 55 60 65 70 75 80 85
56 Temperature (°C)
57
58
59 Fig. S11: Reversibility test of heat engine. Hysteresis is observed for dynamic system in forward
60 and reverse cooling. First heating and then the cooling of the SMA engine was performed.
61
62 35
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 We conducted the study to know the reversibility of our engine. With very slow rate of
5
6 increasing temperature the mechanical behavior (RPM) of the engine evaluated experimentally.
7 As shown in Figure S11 one cycle of gradual heating and cooling followed the different path of
8
9 mechanical behavior. With this experimental results we can conclude that the mention engine is
10 slightly irreversible in nature and doesn’t follow the reversible path.
11
12
13
14
15 (a)
16
17
18
19
20
21
22
23
24
25
26
27
28
Hot
29
30
31 (b)
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Fig. S12: (a) Schematic of multiple shape memory alloy engine in a constant temperature
50
51 reservoir. This experiments helped in understanding the scalability of the technology. (b)
52 Summation of maximum dc power obtained from three different SMA based heat engines.
53
54
55
56
57
58
59
60
61
62 36
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S9. Working principle and thermoelastic cycle of SMA engine:
5
6
7
8
9 (a)
10 1.0
10°C/min
11
12 0.5
13
Heat Flow (W/g)

14
0.0
15
16
17 -0.5
18
19 -1.0
20
21
-1.5
22
23
-80 -60 -40 -20 0 20 40 60 80 100 120
24
25 Temperature (°C)
26
27 (b) (c)
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Fig. S13: (a) DSC analysis of SMA wire with temperature ramping rate of 10°C/min. (b)-(c)
44 Thermodynamic cycle for shape memory alloys [4]. This analysis helped in finding the maximum
45
46 efficiency material.
47
48 Traditionally, the thermodynamic efficiency of a heat engine is obtained using certain idealized
49 thermodynamic cycles, such as Carnot cycle for the ideal heat engine, Joule cycle for the jet
50 turbine engine, and Clausius-Rankine cycle for the steam electric power plant. The one
51
52
fundamental assumption common in all these conceptual cycles is the thermodynamic
53 reversibility of the processes occurring along the thermodynamic paths in these cycles. In
54 reality, none of the actual thermodynamic processes are completely reversible. In the case of
55
56
SMA, the forward and reverse phase transformations are not reversible. Fig. S13 (a) shows the
57 phase change of SMA material with temperature. We used Maxwell model described by
58 Ziolkowski [3] to formulate the thermodynamic cycle of SMA engine. The thermodynamic cycle
59
60
61
62 37
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 of SMA engine is based on the Maxwell model shown in Fig. S13. Fig. S13 (b) is stress-strain
5
6 (σ-ε) diagram and Fig. S13 (c) is temperature-entropy (T-s) diagram[4].
7
8 Thermodynamic efficiency of the heat engine can be shown as
9
10 (s8)
11
12
13 Incase and , equation (s8) reduces to
14
15
16 The SMA element temperature rises from T1 to T2 .
17
18 Tension in the wire causes the stress in the element to increase from to .
19
20 Constant specific heat capacity .
21
22 represents change in entropy due to austenite-martensite phase change and denotes
23 specific heat capacity of the austenite phase.
24
25 S10. Numerical validation of experimental results and cost analysis:
26
27 7
28 Experimental efficiency
29 Numerical TEG efficiency (Tc=22°C)
30 6 SMA Harvester efficiency (Tc=27°C)
31
32
33
34
5
Efficiency (%)

35
36
37
4
38 2.0
39
40 3 1.6
Efficiency (%)

41 1.2
42
43 2 0.8
44 0.4
45
46 1 0.0
55 60 65 70 75 80 85
47 Temperature (°C)
48
49 0
50
51
0 50 100 150 200 250 300
52
53 Temperature (°C)
54 Fig. S14: Comparison of efficiencies between experimental results (Hao et al.[1]) and simulation
55
results of the thermoelectric device for higher temperature (>100°C). Extrapolated simulated
56
57 efficiency results of thermoelectric device for lower temperatures (<100°C) are compared with
58 experimentally obtained SMA-engine.
59
60
61
62 38
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 Cost comparison with TEG
5
6
7
8 In this section we compared our harvesting method with TEG module[5] which operates
9 in the same temperature range (< 85oC). The bulk cost of this device is around $21. This
10 device produced around 3 mW when there is no external air flow. In our case, the cost of
11 1 m of the SMA wire is around $4.75[6]. Currently we are using ~0.5 m of SMA
12
13
material. Fabrication of metallic and plastic pulley could add $3 per heat engine. Thus,
14 the total cost of laboratory scale heat engine will be ~$6. Our heat engine does not
15 require any additional heat sink or forced cooling mechanism, which ensures its hassle-
16 free operation. With the same temperature range of operation, we could generate around
17 18 mW of power, as compared 3 mW from TEG. Thus, our heat engine generates 6 times
18
19 more power at ~1/3rd of the cost compared to TEG.
20
21
22 S11. Setup for water health monitoring:
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Fig. S15: Experimental setup of SMA heat engine for real-time water health monitoring. Low-
42 grade waste thermal energy is converted into mechanical energy and subsequently into electrical
43
energy.
44
45
46
For demonstrating the real applications, we coupled the SMA engine with different kind of
47 devices such as water health monitoring sensors. Fig. S15 shows the experimental setup used for
48 water quality monitoring on the hot water pipes. Setup includes SMA wire, hot pipe, Kapton
49
50 flexible heater (OMEGA Engineering), DC supply (KEITHLEY 2230G-30-1), amplifier, two
51 pulleys, NI-DAQ (M/N: 9171), K-type thermocouple (OMEGA Engineering), step up voltage
52
53
regulator (POLOLU), water quality sensor-TDS-3(HM DIGITAL) and 3D slider.
54
55
56
57
58
59
60
61
62 39
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S12. Dynamic Mechanical Analyzer (DMA) results on wire for 3 days:
5
6
7 100
8 90 90
0.154
9
80
10

Temperature (°C)
Temperature (°C)

11 70
60

Strain (%)
12 60 0.147

13 50
14 30
40
15 0.140
30
16
1100 1200 1300 1400
17 20
Time(minutes)
18 1000 1200 1400
10
0 1000 2000 3000 4000 5000 Time (minutes)
19
20 Time(minutes) 0.156
75
21 70
70 0.154
22
23 65 60
0.152
Stress (MPa)

24
Stress (MPa)

60

Strain (%)
25 50 0.150
26 55
27 0.148
40
28 50
0.146
29 45
30 1000 1200 1400
Time(minutes)
0.144
31 40
32 0 1000 2000 3000 4000 5000 0.142
33 0 1000 2000 3000 4000 5000
Time(minutes)
34 Time (minutes)
35
36
37 Fig. S16: DMA tests for evaluating stress and strain of SMA wire for 3 days. Stress and strain
38 was measured under the closed furnace environment of DMA.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 40
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S13. Brief calculations for carbon foot printing
5
6 If device is mounted directly on pipes
7
8
9
One prototype installed around the circumference of a pipe in vertical position. If we consider
10 the per unit length of pipe, then 1m pipe (if width of device is 5 cm) can easily accommodate 19-
11 20 devices. As demonstrated, 18 mW can be produced with equivalent dimensions by using hot
12
13 water source. Considering the practical operating conditions, it can be safely assumed that 1
14 device of given dimension can produce = 18 mW×0.55 = ~10 mW. Here 0.55 is considered as
15
16
operational factor.
17
18 For 1 m of pipe
19
20 No. of device installed ~20
21
22 In 1 day, the amount of energy produced by 1m of pipe is given by equation s9:
23
24 Energy = Each device power (0.01W) × device (20) × time (24 Hrs.) = 4.8WH (s9)
25
26 where WH=Watt Hours.
27
28 In 3 operating days for which we checked the wire stability under DMA test (Figure S16)), the
29
30 amount of energy produced by 1 m of pipe= 4.8×3 WH=14.4 WH.
31
32 The calculations shown above are for 1m pipe only, but in actual buildings the magnitude of
33 energy generated for 1 day or year will be very high due to availability of long network of pipes.
34
35 If we assume the 500 m of pipe network, then total amount of energy ( E ) for 3 days would be =
36
37 14.4 WH ×500 = 7200 WH =7.2 kWH
38
39 Carbon footprints calculations [7]
40
41 United States Environmental Protection Agency (EPA) reports the emission factor for
42 calculating the CO2 emission saving ( C ) based on electrical energy. This factor is deduced from
43
44 the AVoided Emission and geneRation Tool (AVERT)
45
46 Emission Factor ( EF ) is given as: 7.07 × 10-4 metric tons CO2/kWh. Therefore,
47
48 C  EF  E (s10)
49
50 The expression s10 gives us the carbon saving value of 5.1 kg.
51
52 Our system (in 3 days) in comparison to the energy sources could reduce 5.1 kg of CO2
53
54
55
56
57
58
59
60
61
62 41
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 S14. Long term operation of SMA engine
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 Fig. S17: Long hour of operation of SMA heat engine. Lower pulley shaft rpm is shown for 10
28 hours of operation in lab environment.
29
30 During the heat engine operation, each material point on the wire experiences a cycle of forward
31
and reverse phase transformation within ~1.5 seconds. With this speed, each material point is
32
33 exposed to the hot and cold ambient for approximately 0.0033 and 0.3267 seconds, respectively.
34 By considering the small diameter of the wire (0.44 mm) and moderately low temperature
35 difference between the hot and cold sides, a uniform temperature (and consequently phase
36
37 transformation) distribution in the cross-section can be expected [8, 9]. The observed fatigue
38 resistance of the wires (the wire remains tight and no drop in the efficiency was observed in
39 ~24000 cycles) can be attributed to the uniform distribution of phase transformation in the cross-
40
41 section. The non-uniformity in the phase transformation distribution is one of the main reasons
42 behind initiation and propagation of dislocations, particularly from the grain boundaries in
43 polycrystalline bulk materials [10-12]. The multiple experiments indicated the same efficiency
44
45 and fatigue resistance even for enhanced cycles of operation. Recently, it was reported that by
46 slightly modifying the composition of SMA and modulating phase transition behavior, an
47 extraordinary high fatigue resistance can be obtained [13].
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 42
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1
2
3
4 Supplementary References:
5
6
7
[1] Hao F, Qiu P, Tang Y, Bai S, Xing T, Chu H-S, et al. High efficiency Bi 2 Te 3-based
8 materials and devices for thermoelectric power generation between 100 and 300° C. Energy &
9 Environmental Science. 2016;9:3120-7.
10 [2] Tadesse Y, Thayer N, Priya S. Tailoring the response time of shape memory alloy wires
11 through active cooling and pre-stress. Journal of Intelligent Material Systems and Structures.
12
13 2010;21:19-40.
14 [3] Ziótkowski A. Theoretical analysis of efficiency of shape memory alloy heat engines (based
15 on constitutive models of pseudoelasticity). Mechanics of materials. 1993;16:365-77.
16 [4] Qian S, Ling J, Hwang Y, Radermacher R, Takeuchi I. Thermodynamics cycle analysis and
17
numerical modeling of thermoelastic cooling systems. International Journal of Refrigeration.
18
19 2015;56:65-80.
20 [5] https://www.digikey.com/product-detail/en/marlow-industries,-inc/EHA-PA1AN1-
21 R03/1681-1087-
22 ND/7537072?WT.srch=1&gclid=EAIaIQobChMIpNPlk_Xk3QIVCFSGCh2dOQH7EAQYAiA
23
24
BEgLFzfD_BwE.
25 [6] http://www.dynalloy.com/flexwire_70_90.php.
26 [7]https://www.epa.gov/energy/greenhouse-gases-equivalencies-calculator-calculations-and-
27 references.
28 [8] Mirzaeifar R, DesRoches R, Yavari A. Analysis of the rate-dependent coupled thermo-
29
30 mechanical response of shape memory alloy bars and wires in tension. Continuum Mech Therm.
31 2011;23:363-85.
32 [9] Mirzaeifar R, DesRoches R, Yavari A. Is the Stress Distribution Uniform in the Cross
33 Section of Sma Bars Subjected to Uniaxial Loading? Is It Related to Rate Dependency?
34
35
Proceedings of the Asme Conference on Smart Materials, Adaptive Structures and Intelligent
36 Systems (Smasis 2011), Vol 1. 2012:281-8.
37 [10] Mirzaeifar R, DesRoches R, Yavari A, Gall K. A micromechanical analysis of the coupled
38 thermomechanical superelastic response of textured and untextured polycrystalline NiTi shape
39 memory alloys. Acta Mater. 2013;61:4542-58.
40
41 [11] Yazdandoost F, Mirzaeifar R. Tilt grain boundaries energy and structure in NiTi alloys.
42 Computational Materials Science. 2017;131:108-19.
43 [12] Yazdandoost F, Mirzaeifar R. Generalized stacking fault energy and dislocation properties
44 in NiTi shape memory alloys. Journal of Alloys and Compounds. 2017;709:72-81.
45
[13] Chluba C, Ge W, de Miranda RL, Strobel J, Kienle L, Quandt E, et al. Ultralow-fatigue
46
47 shape memory alloy films. Science. 2015;348:1004-7.
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 43
63
64
65 Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Video
Click here to download Video: Supplementary movie.wmv

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.

You might also like