Next Article in Journal
Synthesis of Novel Polymeric Acrylate-Based Flame Retardants Containing Two Phosphorus Groups in Different Chemical Environments and Their Influence on the Flammability of Poly (Lactic Acid)
Next Article in Special Issue
Enhanced Thermoelectric Performance of Indacenodithiophene-Benzothiadiazole Copolymer Containing Polar Side Chains and Single Wall Carbon Nanotubes Composites
Previous Article in Journal
512-Channel Geometric Droplet-Splitting Microfluidic Device by Injection of Premixed Emulsion for Microsphere Production
Previous Article in Special Issue
Flexible Organic Thermoelectric Materials and Devices for Wearable Green Energy Harvesting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of SrTiO3 Nanoparticles in Conductive Polymer on the Thermoelectric Performance for Efficient Thermoelectrics

School of Chemical Engineering & Materials Science, Chung-Ang University, Seoul 06974, Korea
*
Author to whom correspondence should be addressed.
Polymers 2020, 12(4), 777; https://doi.org/10.3390/polym12040777
Submission received: 18 November 2019 / Revised: 24 March 2020 / Accepted: 27 March 2020 / Published: 1 April 2020
(This article belongs to the Special Issue Polymers for Thermoelectric Applications)

Abstract

:
We present hybrid organic inorganic materials, namely, SrTiO3/polyaniline (PANI) composites, with high thermoelectric performance; samples with various SrTiO3 contents (10, 20, 30, and 50 wt.%) were prepared. The PANI component was obtained through the polymerization of aniline monomers, followed by camphosulfonic acid-doping to enhance its electrical conductivity. SrTiO3, with a high Seebeck coefficient, was used as the N-type inorganic componenet; it was synthesized via a one-pot solvothermal methods and, then, dispersed into the conductive PANI matrix. The SrTiO3 content influenced the Seebeck coefficient and electrical conductivity of the resulting composites. The variations in the thermoelectric properties of the SrTiO3/PANI composites consequently changed their power factor; at room temperature, the highest value was ~49.6 μW·m/K2, which is 17 times larger than that of pure PANI.

1. Introduction

Thermoelectric (TE) power generation is a core technology for renewable energy harvesting and greenhouse gas reduction because of the potential energy conversion between thermal and electrical energies [1,2,3,4,5,6,7,8,9]. The efficiency of TE materials can be evaluated by the dimensionless figure of merit, ZT = S2·σ·T/κ, where S, σ, T, and κ are, respectively, the Seebeck coefficient, the electrical conductivity, the absolute temperature, and the total thermal conductivity. The previous studies generally focused on inorganic TE materials such as Te-based compounds (Bi2Te3, Ag2Te, and Cu2Te) [10,11,12], Se alloys (SnSe, Cu2Se) [13,14,15], and conducting oxides (NaCo2O4, CaMnO3) [16,17]; however, these materials are typically expensive and brittle, which prevents their application in large areas. Thus, polymer-based TE materials have recently been widely used for energy harvesting due to their unique advantages, i.e., low cost, low processing temperature, and mechanical flexibility. In this regard, we have developed high-efficiency metal polymer TE devices [18,19,20].
The ZT can be simply achieved by a low thermal conductivity or with a high power factor (PF = S2·σ). Currently, conductive polymers such as poly(3,4-etylenedioxythiophene)-poly(4-styrenesulfonate) (PEDOT:PSS), polypyrrole, and polythiophene are being used as TE base materials because of their outstanding thermoelectric properties [21,22,23]. Ju et al. enhanced the thermoelectric properties of a conductive PEDOT:PSS matrix by incorporation SnSe nanosheets [24]. SnSe nanosheet/PEDOT:PSS composites have exhibited also enhanced ZT (0.32) at room temperature, which is ~6 times larger than that of pristine PEDOT:PSS. Polyaniline (PANI) is another promising conductive polymer due to its unique characteristics, namely, high σ, and low κ [25,26,27,28]; Anno et al. prepared camphorsulfonic acid (CSA)-doped PANI [29], observing an increased PF of ~0.02 μWm/K2. Wang et al. observed a σ of ~65 S/cm for pure PANI and prepared a PANI/Te nanorod composite via a solution mixing method, achieving high PF (~80 μWm/K2) at high temperature [30].
Perovskite strontium titanate and its composites are recently being widely used as N-type materials due to their high S and chemical stability. In particular, at room temperature, SrTiO3 has an outstanding S (~400 μV/K) [27] compared to other extensively used N-type materials such as Bi2Te3, Sb2Te3, and Ag2Te [31,32,33,34].
In this study, we suggest a strategy to increase the thermoelectric performance by fabrication of hybrid polymer-based TE composites, namely, SrTiO3/PANI composites. The PANI matrix was synthesized through the polymerization of aniline monomers, followed by the doping with CSA to enhance its σ. SrTiO3 was prepared via a one-pot solvothermal method by using Sr and Ti precursor solutions. Then, different SrTiO3 contents were incorporated into the PANI matrix to obtain hybrid organic-inorganic TE composites, whose thermoelectric properties (S, σ, and PF) were successively analyzed. We expected that the SrTiO3 nanoparticles were randomly dispersed within the PANI matrix, providing a highly efficient thermoelectric performance.

2. Experimental

2.1. Materials

Sodium hydroxide (NaOH, 98%), ethanol (C2H5OH, 94%) and hydrochloric acid (HCl, 99%) were purchased from Sigma Aldrich (St. Louis, MO, USA). Strontium nitrate (Sr(NO3)2, 98%), titanium tetraisopropoxide [(CH3)2CHO]4Ti (TTIP, 98%), ammonium persulfate (APS, 98%), aniline (99%), and D(+)-camphorsulfonic acid(CSA, 98%) were provided by Daejung Chemicals & Metals Co. (Seoul, Korea). All chemicals were used without further purification.

2.2. Preparation of PANI

A PANI powder was produced via the chemical polymerization of aniline monomers by using HCl as the dopant and APS as the oxidant. According th the typical synthetic procefure, the monomers (0.05 g) were dispersed in 25 mL of 1M HCl and sonicated the mixture at room temperature. Then, another 1 M HCl solution (1 mL) containing 0.01 g of APS was slowly added to this colloidal solution, followed by steady stirring for 12 h to obtain an aniline/HCL suspension. Next, the polymerization reaction was carried out at 0 °C under stirring for 6 h in an ice bath. The products were filtered and sequentially washed with 1 M HCl, deionized (DI) water, and ethanol, followed by CSA doping to improve the PANI electrical conductivity. The resulting PANI powder was further doped with CSA at a mole ratio of 1:0.6 through a solid-state reaction.

2.3. Synthesis of SrTiO3

First, Sr(NO3)2 (1.9 g) and TTIP (2.55 g) were dissolved in DI water (30 mL) and ethanol (180 mL), respectively. The Sr precursor solution was then slowly added to the Ti one, which was subsequently heated and stirred at 50 °C for 30 min. Next, a 0.2 M NaOH aqueous solution (10 mL) was added to the mixture, followed by stirring for another 30 min. Afterwards, the as-obtained solution was poured into a Teflon-lined autoclave and heated at 140 °C for 24 h. The resulting product (SrTiO3) was filtered, washed repeatedly, and loaded into a quartz ampoule that was successively sealed with a flame at 350 °C for 30 min.

2.4. Preparation of SrTiO3/PANI Composites

A quantity of 0.1 g of the CSA-doped PANI and various amounts (10, 20, 30, and 50 wt.%) of SrTiO3 were dissolved in m-cresol. These mixtures were stirred for 6 h and filtered twice with DI water and ethanol. Finally, the as-prepared SrTiO3/PANI composites were dried in a vacuum oven, ground into fine powders, and hot-pressed at 80 °C for 10 min under 50 MPa.

2.5. Characterization of the SrTiO3/PANI Composites

The morphology of the samples was examined with a field emission scanning electron microscopy (FE-SEM) system (SIGMA, Oberkochen, German), while their element maps were obtained with an energy dispersive X-ray spectroscopy (EDS) instrument (NORAN system 7, Thermo scientific, Seoul, Korea). An X-ray diffraction (XRD) analyzer (New D8-Advance, Bruker-AXS) with Cu Kα radiation (0.154056 nm) was used to identify the crystal phase of the composites; the XRD patterns were acquired in a 2θ range of 20 to 80° at a scan rate of 1°/s. The binding energy peaks were analyzed by using an X-ray photoelectron spectroscopy (XPS) system (Thermo U.K. K-alpha, Seoul, Korea) with Al Kα radiation (1486.6 eV).
Disc-shaped samples with a 12.7 mm diameter were prepared to determine σ and S. The electrical conductivity was derived by using a Keithley 2400 source meter and the four-point probe method; the thickness of the samples was measured with a digital micrometer. The Seebeck coefficient was estimated with homemade equipment consisting of a pair of voltmeters and thermocouples. Five SrTiO3/PANI samples were prepared for the reproducibility of the experiments and the average values are reported here.

3. Results and Discussion

In the formation of SrTiO3 via the one-pot solvothermal method, NaOH plays a key role as follows. The high OH concentration provided by the NaOH solution promotes the hydrolysis of TTIP, yielding negatively charged Ti sol that successively reacts with the Sr cations from the Sr precursor, finally forming SrTiO3 [31].
The XRD patterns of synthesized SrTiO3 (Figure 1a) showed diffraction peaks at 32.4°, 39.9°, 46.5°, 57.8°, 67.8°, and 77.2°. that were attributed to, respectively, the (100), (111), (200), (211), and (220) planes of perovskite SrTiO3 (JCPDS no. 35-0734), indicating the formation of pristine SrTiO3 with a well-defined cubic by structure [35,36].
The successful synthesis and chemical composition of SrTiO3 was then confirmed by the XPS results. In the XPS survey spectrum (Figure 1b), the Sr 3d and Ti 2p peaks were observed. The high-resolution Sr 3d spectrum (Figure 1c) showed peaks at ~132 and ~134 eV that correspond to the Sr 3d5/2 and 3d3/2 binding energies, respectively, of pristine SrTiO3, suggesting its existence in the Sr2+ state. Ti existed as Ti4+, as revealed by the two separate peaks at ~463 and ~458 eV, corresponding to Ti 2p1/2 and Ti 2p3/2, respectively (Figure 1d) [37].
The FE-SEM analysis clearly showed the nano-sized (~50–100 nm) grains of the synthesized SrTiO3 (Figure 2a,b). Furthermore, the corresponding Sr and Ti maps (Figure 2c,d) confirmed its unary composition.
The FT–IR spectra (Figure 3a) exhibited two bands at at ~1580 and ~1450 cm1 and another two at ~1300 and ~1140 cm−1; these signal pairs were attributed to the C=C vibration and the C–N structure, respectively, of the quinoid and benzoid rings in PANI [38,39]. The successful preparation of the SrTiO3/PANI composites as then confirmed by their XRD patterns (Figure 3b), which showed that the SrTiO3 peaks systemically gained intensity as the SrTiO3 content increased. These peaks demonstrate the good dispersion of SrTiO3 and PANI in the sample composite matrix. The FE-SEM observation of the SrTiO3 (30 wt.%)/PANI sample (Figure 3c) further confirmed the uniform dispersion of these two components. Their random dispersion was also demonstrated, along with the successful synthesis of SrTiO3/PANI, via the EDS elemental mapping (Figure 3d–f).
For the measurements of the TE properties, the PANI and SrTiO3/PANI disks were loaded into a Fe mold and pressed at 200 °C under 50 MPa for 10 min.
Figure 4a compared electrical conductivities at room temperature of the various SrTiO3 /PANI samples and pristine PANI (i.e., 0 wt.%). The pristine PANI sample exhibited a σ value of 62.4 S/cm, which is consistent with previous results [26]. As regards the composite sample, σ decreased with the increase in the SrTiO3 contents; this was due to its relatively lower σ of SrTiO3 compared to that of PANI [26]. This trend can be explained by the relation of σ with the charge carrier mobility (μ) and concentration (n): σ = n·e·μ, where e is the electron charge. According to this relation, σ is directly proportional to n and μ, whose values are given in Figure S1. For the SrTiO3/PANI samples, n decreased when increasing the SrTiO3 content, resulting in the above-mentioned σ reduction.
The pristine PANI sample exhibited a positive S (Figure 4b), confirming its P-type semiconducting behavior. On the other hand, the SrTiO3/PANI composites showed negative S values due to the N-type semiconducting behavior of SrTiO3; furthermore, the absolute S value increased along with the SrTiO3 content because of the high S of SrTiO3 at room temperature (~−300 μV/K) [40]. The Sebeck coefficient of a material is generally inversely proportional to μ and this trend is described by the following model [41,42]:
S = 8 · π 2 · k B 2 3 · e · h 2 · m * · T · ( π 3 · n ) 2 3
where kB is the Boltzmann constant, h, is the Planck constant, and m* is the effective mass of the charge carrier.
When increasing the absolute S and decreasing σ, at room temperature, the SrTiO3(20 wt.%)/PANI sample exhibited the largest PF, i.e., ~29.6 μW·m/K2, which is 17 times larger than that of pristine PANI.
Is this study, we prepared SrTiO3/PANI composites and evaluated their thermoelectric properties. With increasing the SrTiO3 content, S accordingly increased while σ decreased. Therefore, the maximum PF was obtained for the composite. These experimental results can provide a strategy for achieving highly efficient TE materials.

4. Conclusions

We fabricate SrTiO3/PANI composites with various SrTiO3 contents and analyzed their thermoelectric properties. The purpose was to fabricate PANI-based materials with high thermoelectric efficiency. The PANI and SrTiO3 components were synthesized via, respectively, the polymerization of aniline monomers and a one-pot solvothermal method with Sr and Ti precursor solutions; NaOH was crucial in the SrTiO3 nanoparticle growth. The morphology and nanostructure of the composites were characterized through XRD, XPS, FT-IR, FE-SEM, and EDS analyses. Upon increasing the SrTiO3 content, σ decreased because of the low σ of SrTiO3, while S increased. Due to these two trends, at room temperature, the sample with 20 wt.% SrTiO3 exhibited the maximum power factor of ~49.6 μW·m/K2, which is ~17 times larger than that of pure PANI. The results of this study demonstrate that the combination of SrTiO3 with PANI as the organic polymer is a successful strategy to fabricate outstanding organic–inorganic TE materials.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/12/4/777/s1, Figure S1: (a) carrier concentration, and (b) Carrier mobility of PANI-SrTiO3 composites with various SrTiO3 contents at room temperature, Figure S2: (a) FE-SEM images of (a). pristine PANI, (b) pristine SrTiO3, and (c) PANI-SrTiO3 composites, Table S1. Seebeck coefficient, electrical conductivity, carrier concentration, and carrier mobility of PANI-SrTiO3 composites at room temperature.

Author Contributions

Conceptualization, J.K.; Data curation, H.J.; Formal analysis, D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the Human Resources Development (No. 20184030202070) of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government Ministry of Trade, Industry and Energy.

Conflicts of Interest

The authors declare no conflict of interest

References

  1. Lin, L.; Wang, M.; Peng, X.; Lissek, E.N.; Mao, Z.; Scarabelli, L.; Adkins, E.; Coskun, S.; Unalan, H.E.; Korgel, B.A.; et al. Opto-thermoelectric nanotweezers. Nat. Photonics 2018, 12, 195–201. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, L.; Chen, Z.-G.; Dargusch, M.S.; Zou, J. High performance thermoelectric Materials: Progress and their applications. Adv. Energy Mater. 2018, 8, 1701797. [Google Scholar] [CrossRef]
  3. Blackburn, J.L.; Ferguson, A.J.; Cho, C.; Grunlan, J.C. Carbon-nanotube-based thermoelectric materials and devices. Adv. Mater. 2018, 30, 1704386. [Google Scholar] [CrossRef] [PubMed]
  4. Yao, H.; Fan, Z.; Cheng, H.; Guan, X.; Wang, C.; Sun, K.; Ouyang, J. Recent development of thermoelectric polymers and composites. Macromol. Rapid Commun. 2018, 39, 1700727. [Google Scholar] [CrossRef] [PubMed]
  5. Snyder, G.J.; Toberer, E.S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. [Google Scholar] [CrossRef]
  6. Zebarjadi, M.; Esfarjani, K.; Dresselhaus, M.; Ren, Z.; Chen, G. Perspectives on thermoelectrics: From fundamentals to device applications. Energy Environ. Sci. 2012, 5, 5147–5162. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, Q.; Sun, Y.; Xu, W.; Zhu, D. Organic thermoelectric materials: Emerging green energy materials converting heat to electricity directly and efficiently. Adv. Mater. 2014, 26, 6829–6851. [Google Scholar] [CrossRef]
  8. Biswas, K.; He, J.; Blum, I.D.; Wu, C.-I.; Hogan, T.P.; Seidman, D.N.; Dravid, V.P.; Kanatzidis, M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 2012, 489, 414. [Google Scholar] [CrossRef]
  9. Liu, W.; Jie, Q.; Kim, H.S.; Ren, Z. Current progress and future challenges in thermoelectric power generation: From materials to devices. Acta Mater. 2015, 87, 357–376. [Google Scholar] [CrossRef] [Green Version]
  10. Fang, H.; Yang, H.; Wu, Y. Thermoelectric properties of silver telluride–bismuth telluride nanowire heterostructure synthesized by site-selective conversion. Chem. Mater. 2014, 26, 3322–3327. [Google Scholar] [CrossRef]
  11. Zhao, W.; Tan, H.T.; Tan, L.P.; Fan, S.; Hng, H.H.; Boey, Y.C.F.; Beloborodov, I.; Yan, Q. N-type carbon nanotubes/silver telluride nanohybrid buckypaper with a high-thermoelectric figure of merit. ACS Appl. Mater. Interfaces 2014, 6, 4940–4946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Park, D.; Ju, H.; Oh, T.; Kim, J. Facile fabrication of one-dimensional Te/Cu2Te nanorod composites with improved thermoelectric power factor and low thermal conductivity. Sci. Rep. 2018, 8, 18082. [Google Scholar] [CrossRef] [Green Version]
  13. Ju, H.; Park, D.; Kim, J. Solution-processable flexible thermoelectric composite films based on conductive polymer/SnSe0.8S0.2 nanosheets/carbon nanotubes for wearable electronic applications. J. Mater. Chem. A 2018, 6, 5627–5634. [Google Scholar] [CrossRef]
  14. Ju, H.; Kim, M.; Park, D.; Kim, J. A strategy for low thermal conductivity and enhanced thermoelectric performance in SnSe: Porous SnSe1–xSx nanosheets. Chem. Mater. 2017, 29, 3228–3236. [Google Scholar] [CrossRef]
  15. Yang, L.; Chen, Z.-G.; Han, G.; Hong, M.; Zou, Y.; Zou, J. High-performance thermoelectric Cu2Se nanoplates through nanostructure engineering. Nano Energy 2015, 16, 367–374. [Google Scholar] [CrossRef]
  16. Terasaki, I. Transport properties and electronic states of the thermoelectric oxide NaCo2O4. Physica B 2003, 328, 63–67. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, Y.; Sui, Y.; Fan, H.; Wang, X.; Su, Y.; Su, W.; Liu, X. High temperature thermoelectric response of electron-doped CaMnO3. Chem. Mater. 2009, 21, 4653–4660. [Google Scholar] [CrossRef]
  18. Ju, H.; Park, D.; Kim, K.; Kim, J. Exfoliated Sn–Se–Te based nanosheets and their flexible thermoelectric composites with poly (3, 4-ethylenedioxythiophene): Poly (styrenesulfonate) fabricated by solution processing. Org. Electron. 2019, 71, 131–135. [Google Scholar] [CrossRef]
  19. Ju, H.; Park, D.; Kim, J. Fabrication of porous SnS nanosheets and their combination with conductive polymer for hybrid thermoelectric application. Chem. Eng. J. 2019, 356, 950–954. [Google Scholar] [CrossRef]
  20. Ju, H.; Park, D.; Kim, J. Conductive polymer based high-performance hybrid thermoelectrics: Polyaniline/tin (II) sulfide nanosheet composites. Polymer 2019, 160, 24–29. [Google Scholar] [CrossRef]
  21. Ju, H.; Kim, M.; Kim, J. Enhanced thermoelectric performance of highly conductive poly (3, 4-ethylenedioxythiophene)/carbon black nanocomposites for energy harvesting. Microelectron. Eng. 2015, 136, 8–14. [Google Scholar] [CrossRef]
  22. Wang, Y.; Yang, J.; Wang, L.; Du, K.; Yin, Q.; Yin, Q.J. Polypyrrole/graphene/polyaniline ternary nanocomposite with high thermoelectric power factor. ACS Appl. Mater. Interfaces 2017, 9, 20124–20131. [Google Scholar] [CrossRef] [PubMed]
  23. Shi, H.; Liu, C.; Xu, J.; Song, H.; Lu, B.; Jiang, F.; Zhou, W.; Zhang, G.; Jiang, Q. Facile fabrication of PEDOT: PSS/polythiophenes bilayered nanofilms on pure organic electrodes and their thermoelectric performance. ACS Appl. Mater. Interfaces 2013, 5, 12811–12819. [Google Scholar] [CrossRef] [PubMed]
  24. Ju, H.; Kim, J. Chemically exfoliated SnSe nanosheets and their SnSe/poly (3, 4-ethylenedioxythiophene): Poly (styrenesulfonate) composite films for polymer based thermoelectric applications. ACS Nano 2016, 10, 5730–5739. [Google Scholar] [CrossRef]
  25. Mitra, M.; Kargupta, K.; Ganguly, S.; Goswami, S.; Banerjee, D. Facile synthesis and thermoelectric properties of aluminum doped zinc oxide/polyaniline (AZO/PANI) hybrid. Synth. Met. 2017, 228, 25–31. [Google Scholar] [CrossRef]
  26. Mitra, M.; Kulsi, C.; Kargupta, K.; Ganguly, S.; Banerjee, D. Composite of polyaniline-bismuth selenide with enhanced thermoelectric performance. J. Appl. Polym. Sci. 2018, 135, 46887. [Google Scholar] [CrossRef]
  27. Chatterjee, M.J.; Banerjee, D.; Chatterjee, K. Composite of single walled carbon nanotube and sulfosalicylic acid doped polyaniline: A thermoelectric material. Mater. Res. Express 2016, 3, 085009. [Google Scholar] [CrossRef]
  28. Wang, W.; Zhang, Q.; Li, J.; Liu, X.; Wang, L.; Zhu, J.; Luo, W.; Jiang, W. An efficient thermoelectric material: Preparation of reduced graphene oxide/polyaniline hybrid composites by cryogenic grinding. RSC Adv. 2015, 5, 8988–8995. [Google Scholar] [CrossRef]
  29. Anno, H.; Hokazono, M.; Akagi, F.; Hojo, M.; Toshima, N. Thermoelectric properties of polyaniline films with different doping concentrations of (±)-10-camphorsulfonic acid. J. Electron. Mater. 2013, 42, 1346–1351. [Google Scholar] [CrossRef]
  30. Wang, Y.; Zhang, S.; Deng, Y. Flexible low-grade energy utilization devices based on high-performance thermoelectric polyaniline/tellurium nanorod hybrid films. J. Mater. Chem. A 2016, 4, 3554–3559. [Google Scholar] [CrossRef]
  31. Liu, D.; Zhang, Y.; Kang, H.; Li, J.; Chen, Z.; Wang, T. Direct preparation of La-doped SrTiO3 thermoelectric materials by mechanical alloying with carbon burial sintering. J. Eur. Ceram. Soc. 2018, 38, 807–811. [Google Scholar] [CrossRef]
  32. Ju, H.; Kim, J. Preparation and structure dependent thermoelectric properties of nanostructured bulk bismuth telluride with graphene. J. Alloys Compd. 2016, 664, 639–647. [Google Scholar] [CrossRef]
  33. Park, D.; Ju, H.; Kim, J. Enhanced thermoelectric power factor and low thermal conductivity in one-dimensional Te/Ag2Te composites. Ceram. Int. 2017, 43, 11156–11162. [Google Scholar] [CrossRef]
  34. Chen, J.; Sun, T.; Sim, D.; Peng, H.; Wang, H.; Fan, S.; Hng, H.H.; Ma, J.; Boey, F.Y.C.; Li, S. Sb2Te3 nanoparticles with enhanced Seebeck coefficient and low thermal conductivity. Chem. Mater. 2010, 22, 3086–3092. [Google Scholar] [CrossRef]
  35. Zhang, Q.; Huang, Y.; Xu, L.; Cao, J.-J.; Ho, W.; Lee, S.C. Visible-light-active plasmonic Ag–SrTiO3 nanocomposites for the degradation of NO in air with high selectivity. ACS Appl. Mater. Interfaces 2016, 8, 4165–4174. [Google Scholar] [CrossRef] [PubMed]
  36. Sharma, D.; Upadhyay, S.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; Dass, S. Improved photoelectrochemical water splitting performance of Cu2O/SrTiO3 heterojunction photoelectrode. J. Phys. Chem. C 2014, 118, 25320–25329. [Google Scholar] [CrossRef]
  37. Xing, G.; Zhao, L.; Sun, T.; Su, Y.; Wang, X. Hydrothermal derived nitrogen doped SrTiO3 for efficient visible light driven photocatalytic reduction of chromium (VI). SpringerPlus 2016, 5, 1132. [Google Scholar] [CrossRef] [Green Version]
  38. Dong, Y.; Zhou, Y.; Ding, Y.; Chu, X.; Wang, C. Sensitive detection of Pb (II) at gold nanoparticle/polyaniline/graphene modified electrode using differential pulse anodic stripping voltammetry. Anal. Methods 2014, 6, 9367–9374. [Google Scholar] [CrossRef]
  39. Kumar, V.; Yokozeki, T.; Goto, T.; Takahashi, T. Synthesis and characterization of PANI-DBSA/DVB composite using roll-milled PANI-DBSA complex. Polymer 2016, 86, 129–137. [Google Scholar] [CrossRef]
  40. Wang, N.; He, H.; Li, X.; Han, L.; Zhang, C. Enhanced thermoelectric properties of Nb-doped SrTiO3 polycrystalline ceramic by titanate nanotube addition. J. Alloys Compd. 2010, 506, 293–296. [Google Scholar] [CrossRef]
  41. Shin, H.S.; Jeon, S.G.; Yu, J.; Kim, Y.-S.; Park, H.M.; Song, J.Y. Twin-driven thermoelectric figure-of-merit enhancement of Bi2Te3 nanowires. Nanoscale 2014, 6, 6158–6165. [Google Scholar] [CrossRef] [PubMed]
  42. Kim, G.H.; Hwang, D.H.; Woo, S.I. Thermoelectric properties of nanocomposite thin films prepared with poly (3, 4-ethylenedioxythiophene) poly (styrenesulfonate) and graphene. Phys. Chem. Chem. Phys. 2012, 14, 3530–3536. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns, along with the XPS (b) survey and high-resolution (c) Sr 3d and (d) Ti 2p spectra, of the synthesized SrTiO3 nanoparticles.
Figure 1. (a) XRD patterns, along with the XPS (b) survey and high-resolution (c) Sr 3d and (d) Ti 2p spectra, of the synthesized SrTiO3 nanoparticles.
Polymers 12 00777 g001
Figure 2. (a,b) Field emission scanning electron microscopy (FE-SEM) and energy dispersive X-ray spectroscopy (EDS) images of SrTiO3. (c,d) Elemental maps of (b), obtained via energy dispersive X-ray spectroscopy.
Figure 2. (a,b) Field emission scanning electron microscopy (FE-SEM) and energy dispersive X-ray spectroscopy (EDS) images of SrTiO3. (c,d) Elemental maps of (b), obtained via energy dispersive X-ray spectroscopy.
Polymers 12 00777 g002
Figure 3. (a) FT–IR spectra of the synthesized polyaniline (PANI), SrTiO3, and SrTiO3 (30 wt.%)/PANI composite. (b) X-ray diffraction (XRD) patterns of pure PANI and the SrTiO3/PANI composites with different contents SrTiO3. (c) FE-SEM images and (df) EDS maps of SrTiO3(30 wt.%)/PANI.
Figure 3. (a) FT–IR spectra of the synthesized polyaniline (PANI), SrTiO3, and SrTiO3 (30 wt.%)/PANI composite. (b) X-ray diffraction (XRD) patterns of pure PANI and the SrTiO3/PANI composites with different contents SrTiO3. (c) FE-SEM images and (df) EDS maps of SrTiO3(30 wt.%)/PANI.
Polymers 12 00777 g003
Figure 4. (a) Electrical conductivity, (b) Seebeck coefficient, and (c) Power factor of the SrTiO3/PANI composites as functions of the SrTiO3 content.
Figure 4. (a) Electrical conductivity, (b) Seebeck coefficient, and (c) Power factor of the SrTiO3/PANI composites as functions of the SrTiO3 content.
Polymers 12 00777 g004

Share and Cite

MDPI and ACS Style

Park, D.; Ju, H.; Kim, J. Effect of SrTiO3 Nanoparticles in Conductive Polymer on the Thermoelectric Performance for Efficient Thermoelectrics. Polymers 2020, 12, 777. https://doi.org/10.3390/polym12040777

AMA Style

Park D, Ju H, Kim J. Effect of SrTiO3 Nanoparticles in Conductive Polymer on the Thermoelectric Performance for Efficient Thermoelectrics. Polymers. 2020; 12(4):777. https://doi.org/10.3390/polym12040777

Chicago/Turabian Style

Park, Dabin, Hyun Ju, and Jooheon Kim. 2020. "Effect of SrTiO3 Nanoparticles in Conductive Polymer on the Thermoelectric Performance for Efficient Thermoelectrics" Polymers 12, no. 4: 777. https://doi.org/10.3390/polym12040777

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop