Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Letter pubs.acs.

org/NanoLett

Flexible Nanocrystal-Coated Glass Fibers for High-Performance Thermoelectric Energy Harvesting


Daxin Liang,, Haoran Yang, Scott W. Finefrock, and Yue Wu*,

State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, People's Republic of China School of Chemical Engineering, Purdue University, West Lafayette, Indiana 47907, United States ABSTRACT: Recent efforts on the development of nanostructured thermoelectric materials from nanowires (Boukai, A. I.; et al. Nature 2008, 451, (7175), 168171; Hochbaum, A. I.; et al. Nature 2008, 451, (7175), 163167) and nanocrystals (Kim, W.; et al. Phys. Rev. Lett. 2006, 96, (4), 045901; Poudel, B.; et al. Science 2008, 320, (5876), 634-638; Scheele, M.; et al. Adv. Funct. Mater. 2009, 19, (21), 34763483; Wang, R. Y.; et al. Nano Lett. 2008, 8, (8), 22832288) show the comparable or superior performance to the bulk crystals possessing the same chemical compositions because of the dramatically reduced thermal conductivity due to phonon scattering at nanoscale surface and interface. Up to date, the majority of the thermoelectric devices made from these inorganic nanostructures are fabricated into rigid configuration. The explorations of truly flexible composite-based flexible thermoelectric devices (See, K. C.; et al. Nano Lett. 2010, 10, (11), 46644667) have thus far achieved much less progress, which in principle could significantly benefit the conversion of waste heat into electricity or the solid-state cooling by applying the devices to any kind of objects with any kind of shapes. Here we report an example using a scalable solution-phase deposition method to coat thermoelectric nanocrystals onto the surface of flexible glass fibers. Our investigation of the thermoelectric properties yields high performance comparable to the state of the art from the bulk crystals and proof-of-concept demonstration also suggests the potential of wrapping the thermoelectric fibers on the industrial pipes to improve the energy efficiency. KEYWORDS: Thermoelectric, flexible, nanocrystal-coated glass fibers, high-performance

ur interest in the flexible thermoelectric fibers origins from its potential impacts in improving energy efficiency from the aspects of power generation, industrial manufacture, and transportation. According to a survey conducted by Lawrence Livermore National Lab in August 2010, nearly 57.8% of the total generated energy in the U.S. in 2009 (54.64 out of 94.57 quads) was rejected into the environment and the majority of it was wasted in the format of heat. Thermoelectric devices, which convert waste heat back to electricity from various sources including residential heating, automotive exhaust, and industrial processes, could significantly improve the energy efficiency and long-term sustainable development of the society. The performance of a thermoelectric material is evaluated by the figure of merit (ZT): ZT = S2T/, where is the electrical conductivity, is the thermal conductivity, S is the Seebeck coefficient, and T is the average temperature of the hot and cold sides of the devices ((Thot + Tcold)/2). However, up to date, most thermoelectric devices are based on rigid yet complicated design, which, together with the high manufacture/installation cost and limited scalability, set a severe barrier for large-scale deployment.17 Thus, a desirable thermoelectric material should not only possess a high power factor (S2) and
XXXX American Chemical Society

a low thermal conductivity but also can be produced through a low-cost yet highly scalable manufacture process. We focus our research on coating the nanocrystals of thermoelectric materials onto the glass fibers, which can be mass-produced in a reel-to-reel manner with nearly unlimited length and has been widely used as thermal insulating material because of its low thermal conductivity (as low as 0.04 W/ mK). We take the advantage of the recent development of colloidal nanocrystal synthesis and use these molecular inks to coat the glass fibers to produce flexible thermoelectric devices.812 As a proof of concept, we chose to work on the PbTe system because of the recent extensive interest in the thermoelectric properties in its bulk alloys and nanostructure forms.1315 In a typical process, we first synthesize PbTe nanocrystals using a modified literature procedure,16 which is a hot injection reaction protected by N2 using a standard Schlenk Line. All chemicals including 1-octadecene (ODE, 90%), oleic acid (OA, 90%), lead(II) oxide (PbO, 99.9+%), tellurium powder (99.8%), hexane (98.5%), acetone (99.5%), hydrazine
Received: February 7, 2012 Revised: March 9, 2012

dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

Nano Letters

Letter

Figure 1. Characterizations of PbTe nanocrystals. (a) XRD diffraction patterns of PbTe nanocrystals (1) before and (2) after annealing at 300 C. Both can be indexed into Altaite PbTe (JCPDS 38-1435). (b) TEM image of PbTe nanocrystals with an average diameter of 13 3 nm. Upper inset shows size distribution of PbTe nanocrystals. (c) HRTEM image of a typical PbTe nanocrystal. Distance between different crystal faces is 0.32 nm, indicating the (200) lattice fringes, which is also the highest peak in XRD. (d) Optical absorption spectrum of PbTe nanocrystals. Inset, the plot of (h)2 versus hshows a band gap of 0.68 eV.

(98%), and acetonitrile (99.8%) were purchased from SigmaAldrich and used without further purification. Tri-n-octylphosphine (TOP, 97%) was purchased from Strem. The TOP-Te solution is prepared by dissolving Te powder in pure TOP in a glovebox with a concentration of 0.75 M and then diluted to 0.5 M with ODE before the injection reaction. In the reaction, 0.223 g of PbO, 0.7 g of OA, and 5 g of ODE are degassed at 140 C for at least 1 h in a 50 mL round-bottom flask under constant N2 flow. Three milliliters of 0.5 M TOP-Te solution is then injected and react at 250 C for 1 min followed by an immediate quench of the reaction by immersing the flask in a water bath. Once the temperature drops to 70 C, 5 mL of hexane is injected and the flask is allowed to cool down to ambient temperature naturally. After that, the product is precipitated and washed with hexane/acetone for 3 times to remove any unreacted precursors. X-ray diffraction (XRD) studies (Figure 1a, red curve) show the materials prepared in this way are Altaite phase PbTe (JCPDS 38-1435). Size distribution observed from lowresolution transmission electron microscopy (TEM) studies (Figure 1b) shows uniform nanocrystals with an average size of 13 3 nm (inset, Figure 1b). High-resolution TEM studies (HRTEM, Figure 1c) confirm that the products are single crystal PbTe nanocrystals and clearly show the (200) lattice fringes, which also agrees well with the fact that (200) is the highest peak observed in XRD pattern for Altaite phase PbTe. Figure 1d shows the optical absorption spectrum of PbTe nanocrystals. The plot of (h)2 (inset, Figure 1d), the square
B

of the absorption coefficient () multiplied by the photon energy (h), versus h (inset, Figure 1d) shows a band gap of 0.68 eV, which is consistent with previous reports on sizedependent bandgap energy observed in colloidal PbTe nanocrystals with various sizes.17 Notably, the band gap energy of our PbTe nanocrystals is larger than the reported bulk value (0.31 eV),17 suggesting the presence of a quantum confinement due to finite size. The coating of glass fibers with PbTe nanocrystals involves the alternative dipping of the glass fibers among PbTe nanocrystal solution, diluted hydrazine aqueous solution, and anhydrous acetonitrile (Figure 2a), which involves the following procedures: (1) Bare fluffy glass fibers are dipped in PbTe nanocrystal solution, then taken out and dried; (2) the fibers are dipped into 0.1 M hydrazine solution (diluted with acetonitrile) to remove the capping ligands on the nanocrystal surface and improve the electrical conductivity;911 (3) the fibers are washed again with 99.8% anhydrous acetonitrile to remove hydrazine and dried in nitrogen flow. This procedure is performed repeatedly until a uniform coating with proper thickness is formed. Furthermore, a thermal annealing in hydrogen/nitrogen mixture (10% H2 diluted in N2) at 300 C for 2 h is performed to improve the mechanical strength of the coating layer on the fibers. Scanning electron microscopy (SEM) studies (Figure 2b) show the coated glass fibers have a uniform diameter of 10 um with 300 nm thick PbTe nanocrystal coating. Notably, through the comparison of the full width at half-maximum of the XRD peaks before (Figure 1a,
dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

Nano Letters

Letter

Figure 2. (a) Scheme of the coating procedure. Insets are pictures of bare glass fibers and PbTe nanocrystal-coated glass fibers, respectively. (b) SEM image of PbTe nanocrystal-coated glass fibers. The diameter of fiber is 10 m. Inset indicates the thickness of PbTe coating is approximately 300 nm. (c) HRTEM image of PbTe nanocrystals after annealing, which is scratched off from the fibers, shows no obvious over growth in nanocrystal size.

red curve) and after the chemical treatment of hydrazine and thermal annealing (Figure 1a, black curve), as well as the HRTEM characterizations of the PbTe layer scratched off from the fibers surface (Figure 2c), no significant change in the size of the PbTe nanocrystals has been observed. However, the chemical treatment and thermal annealing significantly improve the mechanical strength of the coating on the glass fibers and
C

introduce the necking among the nanocrystals to enhance the electrical transport. The glass fibers coated with PbTe nanocrystals have the potential for the thermoelectric application. To explore this exciting opportunity, we have investigated their temperaturedependent electrical conductivity, Seebeck coefficient, thermal conductivity between 300 and 400 K along the axial direction of the fibers, and the typical results are shown in Figure 3. During the measurement, the bundled fibers are fabricated into yarns. The electrical conductivity is measured through a standard four-probe method with a maximum temperature fluctuation of 2 K. The fiber yarns are laid down on measurement substrate and glued on prepatterned gold electrodes using silver paint. The Seebeck coefficient is measured under vacuum by sandwiching the yarn between a heater and heat sink and test the voltage difference between the hot and the cold side with a maximum temperature fluctuation of 0.2 K and a voltage resolution of 50 nV. Thermal conductivity is measured and verified on the yarns using multiple approaches including thermal diffusivity method.18,19 The electrical conductivity (Figure 3a) of the PbTe nanocrystal-coated glass fibers increases from 104.4 Sm1 at 300 K to 172.4 Sm1 at 400 K. Figure 3b shows the temperature dependence of Seebeck coefficient of PbTe nanocrystal-coated glass fibers. The positive Seebeck coefficient value indicates the p-type conduction. The Seebeck coefficient measurement shows an increasing trend from 1201.7 uVK1 at 300 K to 1542.4 uVK1 at 400 K. The thermal conductivity (Figure 3d) at 300 K is measured to be 0.228 Wm1K1 and goes up to 0.234 Wm1K1 around 350 K and then drops to 0.226 Wm1K1 at 400 K. The calculated power factor and ZT for the PbTe nanocrystal-coated glass fibers (Figure 3c,e) increase from 0.15 mWm1K2 and 0.20 at 300 K to 0.41 Wm1K2 and 0.73 at 400 K, respectively. Analysis of these results highlights some important points. First, the electrical conductivity of our PbTe nanocrystal coated thermoelectric fibers is lower than the value of the PbTe bulk crystal (14125 Sm1 at 300 K and 5019 Sm1 at 400 K),20 which is mainly because our PbTe nanocrystal coating layer is only 300 nm thick and the nanocrystal remained in their original size even after the chemical treatment and thermal annealing. The electrical conductivity, in principle, could be further improved through a higher temperature annealing to promote the grain growth and minimize the scattering of charged carriers. However, we decide to keep our annealing process at lower temperature because we noticed in our tests that any prolonged annealing above 400 C could degrade the mechanical stability of the PbTe nanocrystal coating. Second, our PbTe nanocrystal-coated glass fibers show a significantly enhanced Seebeck coefficient than the PbTe bulk crystal within the same temperature range (240 uV/K at 300 K to 350 uV/K at 400 K).21 Conceptually, the significant enhancement of Seebeck coefficient could be explained by quantum confinement and/or energy filtering. Quantum confinement gives rise to delta-function-like density of state (DOS) in zero-dimensional quantum dots, which differs from the DOS in bulk semiconductor with a wide energy distribution and it has been predicted that the delta-function-like DOS will lead to a great enhancement of power factor.22,23 Another possible explanation is through energy filtering, which occurs at graingrain interfaces where charge carriers encounter a potential barrier and only those of them with high enough energy could pass the barrier, which also leads to an increased power factor by theoretical prediction.24 Indeed, our absorption measurement
dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

Nano Letters

Letter

Figure 3. Measurements of thermoelectric properties of PbTe nanocrystal-coated glass fibers. (a) Electrical conductivity, (b) Seebeck coefficient, (c) power factor, (d) thermal conductivity, (e) ZT, and (f) histogram of peak ZT values obtained from different samples.

(Figure 1d) indicates that our PbTe nanocrystals have a much bigger band gap of 0.68 eV compared to the reported bulk value (0.31 eV),17 which suggests that a significant contribution from the quantum confinement may exist. In addition, we have also observed the existing of vast grain boundaries in our samples even after the thermal annealing process (Figure 2c), which also suggests that the energy filtering might play an important role to lead the enhancement in the Seebeck coefficient as well. Third, the thermal conductivity of our PbTe nanocrystal-coated glass fibers (0.228 Wm1K1 at 300 K) is greatly reduced compared to the value from PbTe bulk crystals (2.37 Wm1K1 at 300 K),25 which can be attributed to the fact that the majority of the composite fibers are made from glass and that there is significantly increased phonon scattering at nanocrystal grain boundaries and nanocrystal/glass interfaces. Fourth, the statistic distribution of ZT values (Figure 3F) measured from multiple samples is quite narrow (within 13%), which further proves the uniformity of our nanoparticle coating and device fabrication, providing a reliable and reproducible manufacture route for high-performance thermoelectric devices. The success of functional thermoelectric devices through a simple low-temperature solution coating of nanocrystals on
D

glass fibers suggests the possibility of combining the high performance with other attractive properties including flexibility, lightweight, and shock resistance. As a demonstration, we perform the measurements on the glass fiber yarn coated with PbTe nanocrystals that has bent to a curvature of 84.5 (Figure 4a). The electrical conductivity (Figure 4b) of bent fibers increases from 22.7 Sm1 at 300 K to 53.5 Sm1 at 400 K. Figure 4c shows the temperature dependence of Seebeck coefficient of bent fibers with a decreasing trend from 1100.2 VK1 at 300 K to 1058.0 VK1 at 400 K. The thermal conductivity of bent fibers remains the same without obvious change. The calculated peak power factor and ZT for bent fibers (Figure 4d) at 400 K decrease to 0.060 mWm1K2 and 0.105 as the result of the reduction in electrical conductivity and Seebeck coefficient, which, however, return to the original values of 0.381 mWm1K2 and 0.674 after the bending stress is released. We believe the drop in performance during bending and the full recovery after the bending stress being released are probably due to the micro cracks in the PbTe coating layer, which could open up a wider gap between nanocrystal film domains to lower electrical conductivity when being bent and reseal up after the bending stress is released. In our study, we
dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

Nano Letters

Letter

Figure 4. Measurements of thermoelectric properties of bent PbTe nanocrystal-coated glass fibers. (a) Picture of measurement device with bent fibers. It shows the flexibility of the fibers and the curvature is 84.5. (b) Electrical conductivity and (c) Seebeck coefficient measured from the bent fibers. (d) Comparison of power factor and ZT obtained from bent PbTe nanocrystal-coated glass fibers and the ones released from bending.

Figure 5. Demonstration of using flexible thermoelectric fibers for harvesting energy from industrial pipes. (a) Scheme of the proof-of-concept design. (b) Zero voltage and a stable voltage generated by the thermoelectric fibers without (up panel) and with (lower panel) temperature difference, which could be further optimized through the careful design to reduce thermal interface resistance.

find that the PbTe layer has an extremely stable adhesion on the glass fiber after the annealing process at 300 C for 2 h as described previously. In fact, in order to perform HRTEM studies on the annealed nanocrystals on the fiber (Figure 2c), we had to use a diamond cutter to scratch the surface of the fiber yarn for half an hour to get enough sample. However, the microcracking, although it may not lead to the exfoliation of PbTe nanocrystal coating layer from the glass fibers, could lead to long-term performance stability concern when the fibers are under constant bending/release or at large bending angles. This problem could be minimized by an extra layer of conductive polymers coated on the outside of the PbTe nanocrystal layer, which has been proved to improve the long-term stability of graphene/MnO2 nanostructured electrodes in supercapacitors.26
E

Lastly, we take the first step toward the proof-of-concept demonstration of using the flexible thermoelectric fibers to improve the energy efficiency in industrial process. In our experiment, we simply stuff the PbTe nanocrystal-coated glass fibers between hot and cold water flow through a coaxial design (Figure 5a). The temperature of hot water is 85 C and the temperature of cold water is 27 C. Without turning on the hot water, we do not observe any voltage difference between two electrodes (Figure 5b, up); once the hot water starts to flow in the center tube, an electrical potential of 1.7 mV is established (Figure 5b, down). The voltage here is much smaller than expected simply because of the nonoptimized design with extremely large thermal interface resistance, which significantly reduces the effective temperature difference across the fibers. However, the demonstration here shows a proof of concept to
dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

Nano Letters wrap the thermoelectric fibers onto the industrial pipes, for example, in power plants, chemical/petroleum industry, metal or glass manufacture sites, etc., to act as both thermal insulating materials and recover the waste heat back to electricity to improve energy efficiency. Most importantly, by applying the thin coating of thermoelectric nanocrystals onto glass fibers, significant reduction of raw materials required to build thermoelectric devices can be achieved, dramatically improving the performance/manufacture cost trade-off to enable the largescale installation of thermoelectric energy harvesting devices. In summary, we have demonstrated the ability to rationally fabricate flexible thermoelectric fibers through solution-phase deposition of nanocrystals onto glass fibers. The deposition yields uniform coatings of thermoelectric nanocrystals with precise control over thicknesses. The electrical and thermal transport measurements performed on the fiber devices prove the high quality of the materials yielded from the fabrication and further show their potential applications as high-performance flexible thermoelectric devices for energy harvesting with significantly reduced demand on raw materials toward the economical and scalable conversion of waste heat into electricity, which could be further generalized to other functional thermoelectric nanocrystals, particularly those made from nontoxic and abundant elements.

Letter

AUTHOR INFORMATION

Corresponding Author

*Tel: +1-765-494-6028. E-mail: yuewu@purdue.edu. Fax: +1765-494-0805.


Notes

(11) Ma, W.; Luther, J. M.; Zheng, H.; Wu, Y.; Alivisatos, A. P. Nano Lett. 2009, 9 (4), 16991703. (12) Steinhagen, C.; Panthani, M. G.; Akhavan, V.; Goodfellow, B.; Koo, B.; Korgel, B. A. J. Am. Chem. Soc. 2009, 131 (35), 1255412555. (13) Snyder, G. J.; Toberer, E. S. Nat. Mater. 2008, 7 (2), 105114. (14) Fardy, M.; Hochbaum, A. I.; Goldberger, J.; Zhang, M.; Yang, P. Adv. Mater. 2007, 19 (19), 30473051. (15) Poudeu, P. F. P.; DAngelo, J.; Kong, H.; Downey, A.; Short, J. L.; Pcionek, R.; Hogan, T. P.; Uher, C.; Kanatzidis, M. G. J. Am. Chem. Soc. 2006, 128 (44), 1434714355. (16) Smith, D. K.; Luther, J. M.; Semonin, O. E.; Nozik, A. J.; Beard, M. C. ACS Nano 2011, 5 (1), 183190. (17) Murphy, J. E.; Beard, M. C.; Norman, A. G.; Ahrenkiel, S. P.; Johnson, J. C.; Yu, P. R.; Micic, O. I.; Ellingson, R. J.; Nozik, A. J. J. Am. Chem. Soc. 2006, 128 (10), 32413247. (18) Bhatt, H.; Donaldson, K. Y.; Hasselman, D. P. H; Bhatt, R. T. J. Am. Ceram. Soc. 1990, 73 (2), 312316. (19) Fujishiro, H.; Ikebe, M.; Kashima, T.; Yamanaka, A. Jpn. J. Appl. Phys. 1997, 36 (9A), 56335637. (20) Allgaier, R. S.; Scanlon, W. W. Phys. Rev. 1958, 111 (4), 1029 1037. (21) Heremans, J. P.; Thrush, C. M.; Morelli, D. T. Phys. Rev. B 2004, 70 (11), 115334. (22) Hicks, L. D.; Dresselhaus, M. S. Phys. Rev. B 1993, 47 (19), 1272712731. (23) Hicks, L. D.; Harman, T. C.; Dresselhaus, M. S. Appl. Phys. Lett. 1993, 63 (23), 32303232. (24) Vashaee, D.; Shakouri, A. Phys. Rev. Lett. 2004, 92 (10), 106103. (25) Madelung, O.; Rossler, U.; Schulz, M. Lead telluride (PbTe) thermal conductivity. In SpringerMaterials - The Landolt-Bornstein Database, Collaboration, A; editors of the volumes, I. E. F. C., Eds.; Vol. 41C. (26) Yu, G.; Hu, L.; Liu, N.; Wang, H.; Vosgueritchian, M.; Yang, Y.; Cui, Y.; Bao, Z. Nano Lett. 2011, 11 (10), 44384442.

The authors declare no competing financial interest.

ACKNOWLEDGMENTS D. Liang thanks the support from China Scholarship Council for the exchange study at Purdue University. Y. Wu thanks the support from Purdue University new faculty startup grant, DuPont Young Faculty Award, and a partial support from NSF/DOE Thermoelectric Partnership (Award Number 1048616). REFERENCES

(1) Boukai, A. I.; Bunimovich, Y.; Tahir-Kheli, J.; Yu, J.-K.; Goddard, W. A. III; Heath, J. R. Nature 2008, 451 (7175), 168171. (2) Hochbaum, A. I.; Chen, R.; Delgado, R. D.; Liang, W.; Garnett, E. C.; Najarian, M.; Majumdar, A.; Yang, P. Nature 2008, 451 (7175), 163167. (3) Kim, W.; Zide, J.; Gossard, A.; Klenov, D.; Stemmer, S.; Shakouri, A.; Majumdar, A. Phys. Rev. Lett. 2006, 96 (4), 045901. (4) Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y.; Minnich, A.; Yu, B.; Yan, X.; Wang, D.; Muto, A.; Vashaee, D.; Chen, X.; Liu, J.; Dresselhaus, M. S.; Chen, G.; Ren, Z. Science 2008, 320 (5876), 634638. (5) Scheele, M.; Oeschler, N.; Meier, K.; Kornowski, A.; Klinke, C.; Weller, H. Adv. Funct. Mater. 2009, 19 (21), 34763483. (6) Wang, R. Y.; Feser, J. P.; Lee, J.-S.; Talapin, D. V.; Segalman, R.; Majumdar, A. Nano Lett. 2008, 8 (8), 22832288. (7) See, K. C.; Feser, J. P.; Chen, C. E.; Majumdar, A.; Urban, J. J.; Segalman, R. A. Nano Lett. 2010, 10 (11), 46644667. (8) Guo, Q.; Hillhouse, H. W.; Agrawal, R. J. Am. Chem. Soc. 2009, 131 (33), 1167211673. (9) Luther, J. M.; Beard, M. C.; Song, Q.; Law, M.; Ellingson, R. J.; Nozik, A. J. Nano Lett. 2007, 7 (6), 17791784. (10) Luther, J. M.; Law, M.; Beard, M. C.; Song, Q.; Reese, M. O.; Ellingson, R. J.; Nozik, A. J. Nano Lett. 2008, 8 (10), 34883492.
F
dx.doi.org/10.1021/nl300524j | Nano Lett. XXXX, XXX, XXXXXX

You might also like