Next Article in Journal
A Review on Rhenium Disulfide: Synthesis Approaches, Optical Properties, and Applications in Pulsed Lasers
Next Article in Special Issue
Preparation of a ZnO Nanostructure as the Anode Material Using RF Magnetron Sputtering System
Previous Article in Journal
Green and Scalable Fabrication of Sandwich-like NG/SiOx/NG Homogenous Hybrids for Superior Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Lattice Thermal Conductivity of Transparent Conductive Oxide Thin Film via Localized Vibration of Doping Atoms

1
Department of Materials Science and Engineering, Pusan National University, Busan 46241, Korea
2
Department of Materials Science and Engineering, Ångström Laboratory, Uppsala University, 75321 Uppsala, Sweden
3
Materials Technology and Research, Pusan National University, Busan 46241, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2363; https://doi.org/10.3390/nano11092363
Submission received: 6 August 2021 / Revised: 3 September 2021 / Accepted: 9 September 2021 / Published: 11 September 2021
(This article belongs to the Special Issue Transparent Conductive Nanomaterials: Science and Applications)

Abstract

:
Amorphization using impurity doping is a promising approach to improve the thermoelectric properties of tin-doped indium oxide (ITO) thin films. However, an abnormal phenomenon has been observed where an excessive concentration of doped atoms increases the lattice thermal conductivity (κl). To elucidate this paradox, we propose two hypotheses: (1) metal hydroxide formation due to the low bond enthalpy energy of O and metal atoms and (2) localized vibration due to excessive impurity doping. To verify these hypotheses, we doped ZnO and CeO2, which have low and high bond enthalpies with oxygen, respectively, into the ITO thin film. Regardless of the bond enthalpy energy, the κl values of the two thin films increased due to excessive doping. Fourier transform infrared spectroscopy was conducted to determine the metal hydroxide formation. There was no significant difference in wave absorbance originating from the OH stretching vibration. Therefore, the increase in κl due to the excessive doping was due to the formation of localized regions in the thin film. These results could be valuable for various applications using other transparent conductive oxides and guide the control of the properties of thin films.

Graphical Abstract

1. Introduction

As regulations on carbon emissions are strengthened to prevent global warming, numerous studies of renewable energy are being conducted in various fields to achieve higher energy conversion efficiencies [1,2,3,4,5]. Thermoelectricity is a promising renewable energy source that can replace fossil fuels because it can produce energy through waste heat, unlike other renewable energy sources. To employ thermoelectric (TE) materials in various devices and fields, extensive studies have been carried out to improve the energy conversion efficiency of thin-film thermoelectric materials [6,7,8,9,10,11]. Particularly, as a transparent TE material, tin-doped indium oxide (In2O3:SnO2, ITO) thin film is promising owing to its high electrical conductivity, chemical stability, low toxicity, and low price compared to TE alloys [12,13,14,15]. Accordingly, numerous studies have been carried out to improve the physical properties of ITO, including chemical treatment, impurity doping, and heat treatment, to improve the efficiency compared to opaque TE materials [16,17,18,19,20]. Among them, heat treatment to improve the thin film crystallinity is not promising for ITO-based TE materials because the high crystallinity increases the carrier density (n) of the thin film, which results in a decrease in the Seebeck coefficient (S) and degrades the thermoelectric figure of merit (ZT) owing to the increased thermal conductivity of electrons. On the other hand, a suitable level of impurity doping can induce a high electrical conductivity without increasing n because of the high carrier mobility (μ) [17,21,22]. The electrical conduction of ITO is achieved with a conduction band minimum by overlapping of the In 5s orbital, which is insensitive to metal-oxide bond angle variation in the amorphous structure [23]. Hence, unlike other materials with directional covalent bonds, such as a-Si, the μ of ITO can be maintained in an amorphous thin film by the nondirectional ionic bonding originating from the metal s orbital. Regarding metal oxide thin films, which enable a high μ in an amorphous structure, impurity-doped amorphous ITO thin films are attractive materials for TE thin films because they can have a comparably low n, which decreases the electron thermal conductivity (κe) while maintaining a high electrical conductivity (σ) owing to the high μ.
In our previous study, the microstructure of the ITO thin film was manipulated to control n and thermal conductivity without reducing the electrical conductivity of the thin film by doping with ZnO [24]. The total thermal conductivity (κtot) decreased with the increase in the ZnO concentration. The highest ZT (0.0627) was achieved with an optimized level of ZnO doping. However, an abnormal phenomenon of lattice thermal conductivity (κl) increase was observed when the ZnO concentration exceeded the optimized value, despite the assumption of further progress to a randomly disordered structure proportional to the ZnO concentration. Several studies have been carried out on the increase in κl in relation to the dopant concentration [25,26]. To demonstrate this paradoxical phenomenon, two hypotheses are proposed to explain the κl increase: (1) metal hydroxide formation due to the low bond enthalpy energy of Zn and O and (2) Zn localization due to excessive impurity doping. Further experimental studies are needed to determine the correlation between κl and dopant concentration. To clarify the factor responsible for the increase in κl with a high dopant concentration, an atom that has a high bond enthalpy with O can be employed as a symmetrical dopant.
In this study, we compared the κl values of ZnO- and CeO2-doped ITO, which have low and high bond enthalpies with O, respectively (Zn–O: 159 ± 4 kJ/mol, Ce–O: 795 ± 8 kJ/mol). The ITO thin film was amorphized by doping with ZnO or CeO2 and exhibited a σ and low κl up to the critical level of doped ZnO or CeO2. In addition, we investigated the infrared absorbance using Fourier transform infrared (FTIR) spectroscopy to verify whether this phenomenon is due to the formation of metal hydroxide bonds because of the low bond enthalpy energy between the doped atom and oxygen or formation of localized regions in the amorphous material. There was no remarkable difference in the infrared absorbance of the doped thin film, regardless of the different bond enthalpy energies with oxygen. Thus, the increased lattice thermal conductivity due to the excessive doping of ZnO and CeO2 originates from localized vibration. We expect that these results can be used for other transparent conductive oxide (TCO)-based thin-film systems and significantly contribute to the development of the TCO thin-film field.

2. Materials and Methods

2.1. Thin-Film Fabrication

ITO thin films were deposited on a nonalkali glass using a single sintered ITO target (SnO2: 10 wt %). ITO:Zn and ITO:Ce thin films were deposited by employing ZnO and CeO2 targets and cosputtering with an ITO target (SnO2: 10 wt %) via a magnetron cosputtering system on the nonalkali glass. The base pressure of the deposition process was 1.5 × 10−5 Torr. The total gas pressure was maintained at 1.0 Pa using an Ar gas flow of 20 sccm. All the thin-film samples were deposited to a thickness of 150 nm.

2.2. Thin-Film Characterization

The microstructures of the thin films were analyzed using X-ray diffraction (XRD; D8 Advance, Bruker, Billerica, MA, USA). A Hall effect measurement system (HMS2000, Ecopia, Anyang, Korea) was employed to measure the electrical properties of the thin films. The transmittance was estimated using an ultraviolet (UV)–visible analysis (UV-1800, Shimadzu, Kyoto, Japan). The κtot value of the thin film was measured using the time-domain thermoreflectance (TDTR) method, which uses a 765 nm Ti:sapphire laser. An Al thin film of 85 nm thickness was deposited on the ITO:Zn and an ITO:Ce thin film of 150 nm thickness as the thermal transducer layer of a femtosecond pulsed light beam. The volumetric heat capacity of ITO:Zn, ITO:Ce, and the Al thin film was assumed to be the same as ITO and Al bulk materials [27,28]. The κe value was calculated using the equation κe = L0T/ρ via the carrier density obtained by Hall measurements, where L0, T, and ρ are the Lorentz coefficient, absolute temperature, and resistivity, respectively. After the measurement and calculation, κl was obtained by subtracting κe from κtot. Finally, the OH stretching vibration of the doped ITO thin film was measured using FTIR spectroscopy (Vertex 80v, Bruker, Billerica, MA, USA).

3. Results and Discussion

To analyze the influence of the dopant material in the ITO thin film on the crystallinity of the thin film, ZnO (ITO:Zn) and CeO2 (ITO:Ce) were deposited on a nonalkali glass substrate. As shown in Figure 1, the undoped ITO thin film exhibits the preferred orientations (222) and (400), which indicates that the pure ITO has a polycrystalline structure.
In the case of the doped ITO (ITO:Zn or ITO:Ce), both thin films exhibited amorphous structures regardless of the doping amount. With the increase in the doping amount of ZnO and CeO2, the main peaks of (222) and (400) of ITO disappeared and became broad, which indicates that the thin film transformed the microstructure from polycrystalline to amorphous. Despite the bond enthalpy energy difference of the doped atom with oxygen, the ITO thin film was amorphized even with a small amount of doped ZnO and CeO2.
The electrical properties of the ITO:Zn and ITO:Ce thin films are shown in Figure 2. In the case of ITO:Zn, the lowest resistivity (ρ) was observed for 1 wt % of ZnO owing to the formation of an amorphous-like ternary compound. At concentrations higher than 1 wt %, the n of the ITO:Zn thin film decreased as a function of the ZnO concentration [16,29,30]. In the case of CeO2, the ρ of the ITO:Ce thin film decreased up to 3 wt %; above this concentration, ρ increased. Despite the difference in bond enthalpy with oxygen, both ITO samples exhibited typical electrical property trends in relation to the dopant concentration.
The S value was calculated using the Mott equation with the measured n to analyze the effect of Zn and Ce atoms with different bond enthalpy energies on the thermoelectric performance of the thin film. For metal and degenerated semiconductors, the Mott equation is
S = 8 π 2 k B 2 3 e h 2 ( π 3 n ) 2 3 m * T ,
where kB, h, m*, e, n, and T are the Boltzmann constant, Planck’s constant, effective mass (0.561me = 5.110 × 10−31 kg), electron charge, majority charge carrier density, and absolute temperature, respectively [31]. As shown in Equation (1), the S value is proportional to n−2/3. Figure 3 shows S values plotted in relation to n−2/3, and it shows a linear proportional relationship with respect to n−2/3. The ITO:Zn thin film has a lower n than that of ITO:Ce, which leads to a relatively higher S at a ZnO content of 1 wt %. The highest S was obtained at 9 wt % owing to the considerable decrease in n. With the decrease in n owing to the increased doping amount of CeO2, ITO:Ce exhibited a relatively low S at 9 wt %.
Figure 4 shows the κtot, κe, and κl values of (a) ITO:Zn, (b) ITO:Ce, and (c) ZT of the ITO:Zn and ITO:Ce thin films. The ZT values were calculated using
ZT = σS2T/κtot,
where σ, S, T, and κtot are the electrical conductivity, Seebeck coefficient, absolute temperature, and total thermal conductivity, respectively. The highest ZT (0.0558) was obtained at the highest ZnO dopant concentration owing to the low thermal conductivity. These results indicate relatively high thermoelectric performance in the recently reported TCO-based n-type thermoelectric material [7,9,32,33]. The κl value was obtained by κtot and κe, calculated using κe = L0T/ρ, where L0, T, and ρ are the Lorentz coefficient, absolute temperature, and resistivity, respectively. The κl value can then be obtained using κtot = κl + κe. Regardless of the dopant atom, κtot decreased with the increase in the doping concentration. These trends are affected by the weakened electrical properties and amorphization of the microstructure.
However, an abnormal phenomenon of κl increase above certain dopant concentrations was observed for both ZnO- and CeO2-doped ITO. As mentioned previously, the impurity doping in thin films causes disordered crystallinity, which deteriorates κl. Therefore, κl should also be decreased; however, above a certain concentration, κl is slightly increased. We propose two hypotheses to explain the κl increase: (1) metal hydroxide formation and (2) localization of excessive dopant atoms.
Figure 5 shows the FTIR spectroscopy results for (a) ITO:Zn and (b) ITO:Ce to verify the presence of metal hydroxide in the doped ITO in relation to the dopant material. The FTIR spectroscopy analysis showed wavenumbers of 650 to 4000 cm−1. The peaks at 1023.49–1026.37 cm−1 are related to the phonon mode of the In2O3 lattice. The OH stretching vibration mode is related to the wavenumber of approximately 3240 cm−1 [34,35,36,37,38,39]. The wavenumber region of 650 to 1500 cm−1 corresponds to the typical signal of the ITO thin film, without significant difference, especially near the 1500 cm−1 peak, which originates from O–H bending vibration of absorption water [40,41,42]. In addition, the OH stretching vibration region (wavenumbers around 3240 cm−1) did not show any signals for either ITO:Zn or ITO:Ce thin films despite a high doping concentration of ZnO and CeO2 (9 wt %). This provides clues that the κl increase above the specific dopant concentration is not related to metal hydroxide formation but originates from localized vibration due to the excessive dopant amount. Therefore, the localized region generated by excessive doping of ZnO and CeO2 becomes a new pathway for heat transfer in ITO and suppresses phonon scattering by cations, leading to the increase in κl [25,43].

4. Conclusions

In this study, we prepared ITO:Zn and ITO:Ce thin films with various dopant concentrations to verify the reason for the increase in κl above a certain dopant concentration. The proposed hypotheses were (1) metal hydroxide formation and (2) localized impurity vibration. We employed Zn and Ce, with low and high bond enthalpies with oxygen atoms, respectively. The pure ITO thin film exhibited a polycrystalline structure, while, regardless of the dopant atom, the doped ITO exhibited an amorphous structure. Although there was a difference between the bond enthalpy energies of Zn and Ce with O, the electrical properties of the thin films deteriorated owing to the decrease in n originating from the doping in both ITO:Zn and ITO:Ce. Both ITO:Zn and ITO:Ce exhibited the lowest thermal conductivities at 9 wt % because of the reduced κe due to the significant n reduction. The κl value was increased by excessive doping regardless of the bond enthalpy energy, which indicates that the bond enthalpy does not affect the increase in κl. The reason for the increase in κl was determined by the FTIR spectroscopy results; there was no notable difference in the wavenumber of the OH stretching vibration. The wavenumber region of the OH stretching vibration area did not exhibit any signal, which indicated that the increase in κl did not originate from metal hydroxide formation but from localized vibration with the excessive doping of atoms. We think that these results will guide the future control of the properties of thin films and contribute to next-generation thin-film applications.

Author Contributions

Conceptualization, Y.J.C., S.K., and P.K.S.; methodology, H.Y.L.; validation, Y.J.C., S.K., and P.K.S.; resources, P.K.S.; data curation, Y.J.C., H.Y.L., and S.K.; writing—original draft preparation, Y.J.C.; writing—review and editing, S.K. and P.K.S.; visualization, S.K.; supervision, S.K. and P.K.S.; project administration, P.K.S.; funding acquisition, P.K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Basic Science Research Program through the National Research Foundation of Korea funded by the Ministry of Education (2019R1A6A3A01091664) and partly supported by R&D Platform Establishment of Eco-Friendly Hydrogen Propulsion Ship Program (No. 20006644) and Ministry of Environment (G232019012551). And this research was supported by the BK21 FOUR Program (NO.4120200513801) funded by the Ministry of Education (MOE, Korea) and National Research Foundation of Korea (NRF).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahmoudinezhad, S.; Atouei, S.A.; Cotfas, P.; Cotfas, D.; Rosendahl, L.A.; Rezania, A. Experimental and numerical study on the transient behavior of multi-junction solar cell-thermoelectric generator hybrid system. Energy Convers. Manag. 2019, 184, 448–455. [Google Scholar] [CrossRef]
  2. Park, K.-I.; Xu, S.; Liu, Y.; Hwang, G.-T.; Kang, S.-J.L.; Wang, Z.L.; Lee, K.J. Piezoelectric BaTiO3 thin film nanogenerator on plastic substrates. Nano Lett. 2010, 10, 4939–4943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wu, C.; Wang, A.C.; Ding, W.; Guo, H.; Wang, Z.L. Triboelectric nanogenerator: A foundation of the energy for the new era. Adv. Energy Mater. 2019, 9, 1802906. [Google Scholar] [CrossRef]
  4. Wang, J.; Zhou, S.; Zhang, Z.; Yurchenko, D. High-performance piezoelectric wind energy harvester with Y-shaped attachments. Energy Convers. Manag. 2019, 181, 645–652. [Google Scholar] [CrossRef]
  5. Yarlagadda, V.; Carpenter, M.K.; Moylan, T.E.; Kukreja, R.S.; Koestner, R.; Gu, W.; Thompson, L.; Kongkanand, A. Boosting fuel cell performance with accessible carbon mesopores. ACS Energy Lett. 2018, 3, 618–621. [Google Scholar] [CrossRef] [Green Version]
  6. Juang, Z.-Y.; Tseng, C.-C.; Shi, Y.; Hsieh, W.-P.; Ryuzaki, S.; Saito, N.; Hsiung, C.-E.; Chang, W.-H.; Hernandez, Y.; Han, Y.; et al. Graphene-Au nanoparticle based vertical heterostructures: A novel route towards high-ZT Thermoelectric devices. Nano Energy 2017, 38, 385–391. [Google Scholar] [CrossRef] [Green Version]
  7. Liu, S.; Lan, M.; Li, G.; Piao, Y.; Ahmoum, H.; Wang, Q. Breaking the tradeoff among thermoelectric parameters by multi composite of porosity and CNT in AZO films. Energy 2021, 225, 120320. [Google Scholar] [CrossRef]
  8. Liu, S.; Li, G.; Lan, M.; Zhu, M.; Mori, T.; Wang, Q. Improvement of Thermoelectric Properties of Evaporated ZnO:Al Films by CNT and Au Nanocomposites. J. Phys. Chem. C 2020, 124, 12713–12722. [Google Scholar] [CrossRef]
  9. Tambasov, I.A.; Voronin, A.S.; Evsevskaya, N.P.; Volochaev, M.N.; Fadeev, Y.V.; Simunin, M.M.; Aleksandrovsky, A.S.; Smolyarova, T.E.; Abelian, S.R.; Tambasova, E.V.; et al. Thermoelectric properties of low-cost transparent single wall carbon nanotube thin films obtained by vacuum filtration. Phys. E 2019, 114, 113619. [Google Scholar] [CrossRef]
  10. Tomeda, A.; Ishibe, T.; Taniguchi, T.; Okuhata, R.; Watanabe, K.; Nakamura, Y. Enhanced thermoelectric performance of Ga-doped ZnO film by controlling crystal quality for transparent thermoelectric films. Thin Solid Films 2018, 666, 185–190. [Google Scholar] [CrossRef]
  11. Yang, C.; Souchay, D.; Kneiß, M.; Bogner, M.; Wei, H.M.; Lorenz, M.; Oeckler, O.; Benstetter, G.; Fu, Y.Q.; Grundmann, M. Transparent flexible thermoelectric material based on non-toxic earth-abundant p-type copper iodide thin film. Nat. Commun. 2017, 8, 16076. [Google Scholar] [CrossRef] [PubMed]
  12. Bell, L.E. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hong, J.-E.; Lee, S.-K.; Yoon, S.-G. Enhanced thermoelectric properties of thermal treated Sb2Te3 thin films. J. Alloys Compd. 2014, 583, 111–115. [Google Scholar] [CrossRef]
  14. Jeong, M.-W.; Na, S.; Shin, H.; Park, H.-B.; Lee, H.-J.; Joo, Y.-C. Thermomechanical in situ monitoring of Bi2Te3 thin film and its relationship with microstructure and thermoelectric performances. Electron. Mater. Lett. 2018, 14, 426–431. [Google Scholar] [CrossRef]
  15. Loureiro, J.; Neves, N.; Barros, R.; Mateus, T.; Santos, R.; Filonovich, S.; Reparaz, S.; Sotomayor-Torres, C.M.; Wyczisk, F.; Divay, L. Transparent aluminium zinc oxide thin films with enhanced thermoelectric properties. J. Mater. Chem. A 2014, 2, 6649–6655. [Google Scholar] [CrossRef]
  16. Ali, H. Characterization of a new transparent-conducting material of ZnO doped ITO thin films. Phys. Status Solidi A 2005, 202, 2742–2752. [Google Scholar] [CrossRef]
  17. Chung, S.M.; Shin, J.H.; Cheong, W.-S.; Hwang, C.-S.; Cho, K.I.; Kim, Y.J. Characteristics of Ti-doped ITO films grown by DC magnetron sputtering. Ceram. Int. 2012, 38, S617–S621. [Google Scholar] [CrossRef]
  18. Fallah, H.R.; Ghasemi, M.; Hassanzadeh, A. Influence of heat treatment on structural, electrical, impedance and optical properties of nanocrystalline ITO films grown on glass at room temperature prepared by electron beam evaporation. Phys. E 2007, 39, 69–74. [Google Scholar] [CrossRef]
  19. Hu, Y.; Diao, X.; Wang, C.; Hao, W.; Wang, T. Effects of heat treatment on properties of ITO films prepared by rf magnetron sputtering. Vacuum 2004, 75, 183–188. [Google Scholar] [CrossRef]
  20. Wu, C.; Wu, C.; Sturm, J.; Kahn, A. Surface modification of indium tin oxide by plasma treatment: An effective method to improve the efficiency, brightness, and reliability of organic light emitting devices. Appl. Phys. Lett. 1997, 70, 1348–1350. [Google Scholar] [CrossRef] [Green Version]
  21. Yang, C.-H.; Lee, S.-C.; Lin, T.-C.; Zhuang, W.-Y. Opto-electronic properties of titanium-doped indium–tin-oxide films deposited by RF magnetron sputtering at room temperature. Mat. Sci. Eng. B-Adv. 2006, 134, 68–75. [Google Scholar] [CrossRef]
  22. Kang, Y.; Kwon, S.; Choi, J.; Cho, Y.; Song, P. Properties of Ce-doped ITO films deposited on polymer substrate by DC magnetron sputtering. Thin Solid Films 2010, 518, 3081–3084. [Google Scholar] [CrossRef]
  23. Nomura, K.; Ohta, H.; Takagi, A.; Kamiya, T.; Hirano, M.; Hosono, H. Room-temperature fabrication of transparent flexible thin-film transistors using amorphous oxide semiconductors. Nature 2004, 432, 488–492. [Google Scholar] [CrossRef]
  24. Lee, H.Y.; Yang, I.J.; Yoon, J.-H.; Jin, S.-H.; Kim, S.; Song, P.K. Thermoelectric Properties of Zinc-Doped Indium Tin Oxide Thin Films Prepared Using the Magnetron Co-Sputtering Method. Coatings 2019, 9, 788. [Google Scholar] [CrossRef] [Green Version]
  25. Cocemasov, A.; Brinzari, V.; Jeong, D.-G.; Korotcenkov, G.; Vatavu, S.; Lee, J.-S.; Nika, D.L. Thermal transport evolution due to nanostructural transformations in Ga-doped indium-tin-oxide thin films. Nanomaterials 2021, 11, 1126. [Google Scholar] [CrossRef]
  26. Lan, J.-L.; Liu, Y.; Lin, Y.-H.; Nan, C.-W.; Cai, Q.; Yang, X. Enhanced thermoelectric performance of In2O3-based ceramics via Nanostructuring and Point Defect Engineering. Sci. Rep. 2015, 5, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Lide, D.R. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2004; Volume 85. [Google Scholar]
  28. Yagi, T.; Tamano, K.; Sato, Y.; Taketoshi, N.; Baba, T.; Shigesato, Y. Analysis on thermal properties of tin doped indium oxide films by picosecond thermoreflectance measurement. J. Vac. Sci. Technol. A 2005, 23, 1180–1186. [Google Scholar] [CrossRef]
  29. Carreras, P.; Antony, A.; Roldán, R.; Nos, O.; Frigeri, P.A.; Asensi, J.M.; Bertomeu, J. Transparent conducting thin films by co-sputtering of ZnO-ITO targets. Phys. Status Solidi C 2010, 7, 953–956. [Google Scholar] [CrossRef] [Green Version]
  30. Liu, D.-S.; Wu, C.-C.; Lee, C.-T. A transparent and conductive film prepared by RF magnetron cosputtering system at room temperature. Jpn. J. Appl. Phys. 2005, 44, 5119. [Google Scholar] [CrossRef]
  31. Snyder, G.J.; Toberer, E.S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. [Google Scholar] [CrossRef] [PubMed]
  32. Ishibe, T.; Tomeda, A.; Komatsubara, Y.; Kitaura, R.; Uenuma, M.; Uraoka, Y.; Yamashita, Y.; Nakamura, Y. Carrier and phonon transport control by domain engineering for high-performance transparent thin film thermoelectric generator. Appl. Phys. Lett. 2021, 118, 151601. [Google Scholar] [CrossRef]
  33. Liu, S.; Li, G.; Lan, M.; Zhu, M.; Miyazaki, K.; Wang, Q. Role of intrinsic defects on thermoelectric properties of ZnO:Al films. Ceram. Int. 2021, 47, 17760–17767. [Google Scholar] [CrossRef]
  34. Petit, T.; Puskar, L. FTIR spectroscopy of nanodiamonds: Methods and interpretation. Diam. Relat. Mater. 2018, 89, 52–66. [Google Scholar] [CrossRef]
  35. Winiarski, J.; Tylus, W.; Winiarska, K.; Szczygieł, I.; Szczygieł, B. XPS and FT-IR characterization of selected synthetic corrosion products of zinc expected in neutral environment containing chloride ions. J. Spectrosc. 2018, 2018, 2079278. [Google Scholar] [CrossRef] [Green Version]
  36. Wu, N.C.; Shi, E.W.; Zheng, Y.Q.; Li, W.J. Effect of pH of medium on hydrothermal synthesis of nanocrystalline cerium (IV) oxide powders. J. Am. Ceram. Soc. 2002, 85, 2462–2468. [Google Scholar] [CrossRef]
  37. Zamiri, R.; Abbastabar Ahangar, H.; Kaushal, A.; Zakaria, A.; Zamiri, G.; Tobaldi, D.; Ferreira, J. Dielectrical properties of CeO2 nanoparticles at different temperatures. PLoS ONE 2015, 10, e0122989. [Google Scholar]
  38. Petrov, T.; Markova-Deneva, I.; Chauvet, O.; Nikolov, R.; Denev, I. Sem and Ft-Ir Spectroscopy Study of Cu, Sn AND Cu-Sn Nanoparticles. J. Chem. Technol. Metall. 2012, 47, 2. [Google Scholar]
  39. Zhu, G.; Guo, L.; Shen, X.; Ji, Z.; Chen, K.; Zhou, H. Monodispersed In2O3 mesoporous nanospheres: One-step facile synthesis and the improved gas-sensing performance. Sens. Actuators B Chem. 2015, 220, 977–985. [Google Scholar] [CrossRef]
  40. Chang, K.-C.; Tsai, T.-M.; Chang, T.-C.; Zhang, R.; Chen, K.-H.; Chen, J.-H.; Chen, M.-C.; Huang, H.-C.; Zhang, W.; Lin, C.-Y. Improvement of resistive switching characteristic in silicon oxide-based RRAM through hydride-oxidation on indium tin oxide electrode by supercritical CO2 fluid. IEEE Electron. Device Lett. 2015, 36, 558–560. [Google Scholar] [CrossRef]
  41. Raja, K.; Ramesh, P.; Geetha, D. Structural, FTIR and photoluminescence studies of Fe doped ZnO nanopowder by co-precipitation method. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 131, 183–188. [Google Scholar] [CrossRef] [PubMed]
  42. Ali, A.; Ansari, A.A.; Kaushik, A.; Solanki, P.R.; Barik, A.; Pandey, M.; Malhotra, B. Nanostructured zinc oxide film for urea sensor. Mater. Lett. 2009, 63, 2473–2475. [Google Scholar] [CrossRef]
  43. Liu, W.; Shi, X.; Hong, M.; Yang, L.; Moshwan, R.; Chen, Z.-G.; Zou, J. Ag doping induced abnormal lattice thermal conductivity in Cu2Se. J. Mater. Chem. C 2018, 6, 13225–13231. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) ZnO-doped ITO and (b) CeO2-doped ITO. The pure ITO exhibits a polycrystalline structure. The crystal structure of ITO doped with ZnO and CeO2 becomes amorphous even with a small level of doping. There is no change in the diffraction pattern when the doping amount is further increased.
Figure 1. XRD patterns of (a) ZnO-doped ITO and (b) CeO2-doped ITO. The pure ITO exhibits a polycrystalline structure. The crystal structure of ITO doped with ZnO and CeO2 becomes amorphous even with a small level of doping. There is no change in the diffraction pattern when the doping amount is further increased.
Nanomaterials 11 02363 g001
Figure 2. Comparison of electrical properties of (a) ZnO-doped ITO and (b) CeO2-doped ITO depending on the concentration of doping atoms. With a smaller amount of doping, the resistivities of both thin films decrease. On the other hand, excessive doping degrades the electrical properties of the thin film.
Figure 2. Comparison of electrical properties of (a) ZnO-doped ITO and (b) CeO2-doped ITO depending on the concentration of doping atoms. With a smaller amount of doping, the resistivities of both thin films decrease. On the other hand, excessive doping degrades the electrical properties of the thin film.
Nanomaterials 11 02363 g002
Figure 3. Seebeck coefficients (S) and n−(2/3) of (a) ZnO-doped ITO and (b) CeO2-doped ITO at various dopant concentrations. The Mott equation for metals and degenerated semiconductors was employed to calculate S using the measured carrier density.
Figure 3. Seebeck coefficients (S) and n−(2/3) of (a) ZnO-doped ITO and (b) CeO2-doped ITO at various dopant concentrations. The Mott equation for metals and degenerated semiconductors was employed to calculate S using the measured carrier density.
Nanomaterials 11 02363 g003
Figure 4. Dependence of the thermal conductivities of the (a) ZnO-doped ITO and (b) CeO2-doped ITO on the doping concentration. (c) Thermoelectric figures of merit (ZT) of ZnO- and CeO2-doped ITO. The lattice thermal conductivity increased above a certain doping concentration.
Figure 4. Dependence of the thermal conductivities of the (a) ZnO-doped ITO and (b) CeO2-doped ITO on the doping concentration. (c) Thermoelectric figures of merit (ZT) of ZnO- and CeO2-doped ITO. The lattice thermal conductivity increased above a certain doping concentration.
Nanomaterials 11 02363 g004
Figure 5. FTIR spectroscopy of (a) ZnO-doped ITO and (b) CeO2-doped ITO. The inset shows a detail of the OH stretching vibration area. Although Zn and Ce have different bond enthalpies with oxygen, there is no change in the OH stretching vibration region according to the doping amount.
Figure 5. FTIR spectroscopy of (a) ZnO-doped ITO and (b) CeO2-doped ITO. The inset shows a detail of the OH stretching vibration area. Although Zn and Ce have different bond enthalpies with oxygen, there is no change in the OH stretching vibration region according to the doping amount.
Nanomaterials 11 02363 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Choi, Y.J.; Lee, H.Y.; Kim, S.; Song, P.K. Controlled Lattice Thermal Conductivity of Transparent Conductive Oxide Thin Film via Localized Vibration of Doping Atoms. Nanomaterials 2021, 11, 2363. https://doi.org/10.3390/nano11092363

AMA Style

Choi YJ, Lee HY, Kim S, Song PK. Controlled Lattice Thermal Conductivity of Transparent Conductive Oxide Thin Film via Localized Vibration of Doping Atoms. Nanomaterials. 2021; 11(9):2363. https://doi.org/10.3390/nano11092363

Chicago/Turabian Style

Choi, Young Joong, Ho Yun Lee, Seohan Kim, and Pung Keun Song. 2021. "Controlled Lattice Thermal Conductivity of Transparent Conductive Oxide Thin Film via Localized Vibration of Doping Atoms" Nanomaterials 11, no. 9: 2363. https://doi.org/10.3390/nano11092363

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