Enhanced thermoelectric properties of p-type polycrystalline SnSe by regulating the anisotropic crystal growth and Sn vacancy
Liu Chengyan1, Miao Lei1, †, Wang Xiaoyang1, Wu Shaohai1, Zheng Yanyan1, Deng Ziyang1, Chen Yulian1, Wang Guiwen2, Zhou Xiaoyuan2
Guangxi Key Laboratory of Information Material, Guangxi Collaborative Innovation Center of Structure and Property for New Energy and Materials, School of Material Science and Engineering,Guilin University of Electronic Technology, Guilin 541004, China
College of Physics, Chongqing University, Chongqing 401331, China

 

† Corresponding author. E-mail: miaolei@guet.edu.cn

Abstract

Thermoelectric selenides have attracted more and more attentions recently. Herein, p-type SnSe polycrystalline bulk materials with good thermoelectric properties are presented. By using the SnSe2 nanostructures synthesized via a wet-chemistry route as the precursor, polycrystalline SnSe bulk materials were successfully obtained by a combined heat-treating process under reducing atmosphere and following spark plasma sintering procedure. As a reference, the SnSe nanostructures synthesized via a wet-chemistry route were also fabricated into polycrystalline bulk materials through the same process. The thermoelectric properties of the SnSe polycrystalline transformed from SnSe2 nanostructures indicate that the increasing of heattreating temperature could effectively decrease the electrical resistivity, whereas the decrease in Seebeck coefficient is nearly invisible. As a result, the maximum power factor is enhanced from to at 612 °C. On the other hand, the reference sample, which was obtained by using SnSe nanostructures as the precursor, displays very poor power factor of only at 537 °C. The x-ray diffraction (XRD), scanning electron microscope (SEM), x-ray fluorescence (XRF), and Hall effect characterizations suggest that the anisotropic crystal growth and existing Sn vacancy might be responsible for the enhanced electrical transport in the polycrystalline SnSe prepared by using SnSe2 precursor. On the other hand, the impact of heat-treating temperature on thermal conductivity is not obvious. Owing to the boosting of power factor, a high zT value of 1.07 at 612 °C is achieved. This study provides a new method to synthesize polycrystalline SnSe and pave a way to improve the thermoelectric properties of polycrystalline bulk materials with similar layered structure.

1. Introduction

Thermoelectric materials are becoming more and more concerned, aiming at boosting their applications in waste-heat recovery and cooling.[1,2] Generally, the thermoelectric conversion efficiency can be determined by a figure of merit , where T, S, σ, and κ are absolute temperature, the Seebeck coefficient, electrical conductivity, and thermal conductivity of a thermoelectric material, respectively. Obviously, high zT is equivalent to high power factor ( and low thermal conductivity simultaneously. In the past decade, several strategies were usually adopted to enhance the power factor of bulk materials, such as carrier concentration engineering,[3,4] increase of band degeneracy,[46] enhancement of band effective mass,[7,8] and improvement of carrier mobility.[9] To decrease the thermal conductivity, which contributes most significant advances in improving zT, there are several principles, including introducing point defects,[10] second-phase nanostructures,[1113] all-length-scale hierarchical architectures,[14,15] and choosing compounds with complex crystal structures.[16]

In 2014, Zhao et al. reported that a record zT ∼2.6 at 923 K was achieved along the b axis of SnSe single crystal, benefiting from the low lattice thermal conductivity ( at 973 K) which was mainly ascribed to large phonon anharmonicity.[17,18] Furthermore, enhanced power factor and ultrahigh average zT (1.60) in the range from 300 K to 923 K were realized in Na-doped samples.[4,6,19] These exciting phenomena have stimulated much interest on this selenide which consists of earth-abundant and environmentally friendly elements. However, the application of thermoelectric SnSe single crystals is hindered because their growth is challenging due to a strong phase transition at high temperature and machinability is poor owing to the weak chemical bonds along a axis. As a result, considerable effort has been focused on improving the thermoelectric performance of polycrystalline SnSe.[2025] Unfortunately, up to date, the ultrahigh zT values of SnSe crystals have not been reproduced in polycrystalline SnSe because polycrystalline samples show lower carrier mobility and higher thermal conductivity. It is reasonable to suppose that the carrier scattering at grain boundaries and random distribution of crystalline grains with strong anisotropy should decrease the carrier mobility. Unexpectedly, the higher lattice thermal conductivity of polycrystalline SnSe is reported by lots of papers, which is at odds with the simple expectation that grain boundary scattering should act as an additional source of phonon scattering.[26] The possible oxidation in polycrystalline SnSe could reconcile this contradicting result since SnO has very high thermal conductivity ( ).[27,28] Besides, vast off-stoichiometric defects[10] and relatively low density[29] may provide additional contribution to the ultralow lattice thermal conductivity.

To improve the thermoelectric performance of polycrystalline SnSe, Na,[30] K,[21,31] Ag,[22] Br,[23] I,[24] and BiCl3[25] have all been proven to be effective dopants in both p-type and n-type samples. Nevertheless, the maximum zT nearly keeps at relatively low values ( ) though the carrier concentration is effectively increased. SnSe crystallizes in a highly anisotropic layered orthorhombic (Pnma group) crystal structure below 800 K and undergoes a structural transition from Pnma to Cmcm above 800 K.[32,33] The layered structure suggests that the control of crystalline orientation is very significant in polycrystalline SnSe samples.[34,35] Besides, Sn vacancy always plays an important role in improving the thermoelectric properties because it resembles hole-doping which can enhance the carrier concentration of p-type SnSe by shifting the Fermi level and simultaneously inducing localized flat bands in the vicinity of the Fermi energy level.[3638] Meanwhile, it also acts as point defect to scatter phonon and then decrease the thermal conductivity. These motivate us to optimize the thermoelectric properties by simultaneously regulating the anisotropic crystal growth and Sn vacancy.

In this study, the SnSe polycrystalline bulk materials were successfully prepared by heat treating the SnSe2 nanostructures synthesized via a wet-chemistry route at temperature above 600 °C. SnSe2 is a semiconductor with layered CdI2-type structure, and has very strong anisotropic electrical and thermal properties.[3942] This distinctive crystalline orientation of precursor SnSe2 may affect the anisotropic crystal growth during the transformation of SnSe2 to SnSe. Furthermore, vast Sn vacancy should be formed during this process because of the sublimation of atoms. In fact, the measurement of thermoelectric properties demonstrate that the intensified anisotropic crystal growth and Sn vacancy enhance the power factor from to at 612 °C. Then, a high zT value of 1.07 at 612 °C is obtained.

2. Experimental section
2.1. Chemical reagents

Tin chloride dihydrate ( , 99.99%), selenium dioxide (SeO2), and hydrazine hydrate ( , 98%) were purchased from Aladdin industrial corporation, Shanghai, China. All the chemicals were analytical grade and used without purification.

2.2. Synthesis of SnSe2 and SnSe nanostructures

The SnSe2 and SnSe nanostructures were synthesized by a condensing reflux method. In a typical synthesis process of SnSe2 nanostructures, 2.256 g of and 2.220g of SeO2 were added into 250 ml of glycol with magnetic stirring in a 500-ml round bottom flask with three necks, which was heated by an oil-bath. Then 6-mL of was poured into the above solution when the temperature in glycol reaches 80 °C (Attention! A large amount of heat and gas (or foam) generated as soon as the hydrazine was added). After further increasing to the assigned temperature, the status of flask was maintained for 24 h in the oil-bath. In this experiment, the highest temperature in glycol would reach 182 °C (The solution was boiling). After the reaction, it was cooled down to room temperature naturally. The black powders were collected after washing using absolute ethanol and deionized water, and then dried under vacuum at 60 °C for 12 h. SnSe nanostructures were synthesized by the same process except that the molar ratio of Sn to Se was 1:1, namely 2.256 g of and 1.110 g of SeO2.

2.3. Fabrication of the SnSe polycrystalline bulk materials

At first, the SnSe2 and SnSe nanostructures synthesized via a wet-chemistry route were cold-pressed into cylinders with a diameter of 20 mm under a force of 9.8 kN. After the heat-treating process under mixed gases (8% H2 + 92% Ar) for protecting the samples from oxidation at different temperatures for 15 h in a tubular furnace, the cylinders were grounded into powders again. The flow rate of protective gases was 0.15 L/min and the heating rate of tubular furnace was 5 °C/min. Then, the powders were fabricated into cylinders with a diameter of 20 mm again by a spark plasma sintering (SPS, LABOX-325, SINTER LAND INC.) method with a pressure of 50 MPa at 360 °C (for the samples whose heat-treating temperature was 400 °C and 500 °C) and 500 °C (for the other samples) for 6 min. The sintered samples from SnSe2 nanostructures are identified by T@SnSe2, where T is heat-treating temperature, while the samples from SnSe nanostructures corresponding to T@SnSe. Finally, prisms with the size of about 3 mm × 3 mm × 15 mm were cut from the sintered cylinders for the characterization of electrical resistivity and Seebeck coefficient, and wafers with a diameter of 10 mm and 5 mm were prepared for the measurement of thermal diffusivity and specific heat, respectively.

2.4. Characterization

The crystal structures of the powders and bulk materials were characterized by powder diffractometry (PANalytical X’pert Pro MPD operated at 40 kV and 40 mA, CuKa, λ = 0.154 nm). And their morphology observation was performed on a field-emission scanning-electron microscopy (FESEM; Hitachi S-4800, Japan) and transmission electron microscopy (TEM) (Tecnai G220). The composition was determined by x-ray fluorescence (XRF) analysis. Electrical resistivity ( and Seebeck coefficient (S) were measured by the static DC method (ZEM-3; ULVAC RIKO, Japan) in a low-helium (99.999%) atmosphere with a temperature gradient from 20 °C to 30 °C and 40 °C. The thermal conductivity ( is calculated from , where D is the thermal diffusivity, is the specific heat, and ρ is the measured density. The thermal diffusivity (D) was measured by a laser flash method (LFA-457; NETZSCH); the specific heat was determined by differential scanning calorimetry (DSC) (DSC-404F3, NETZSCH); and the volume density ( was determined by Archimedes’ method. The Hall coefficient ( at room temperature was measured in a Quantum Design, Inc. Physical Properties Measurement System (PPMS). The Hall carrier concentration (n) and mobility (μ) were evaluated by using the formula and , respectively.

3. Results and discussion

In this experiments, the temperature of the condensing reflux method in solvent was regulated. After the reaction, black powders were obtained. The corresponding XRD patterns are shown in Fig. 1(a). From this figure, we could conclude that only peaks well-matched with Se (JCPDS 01-086-2246) are observed if the temperature does not exceed 100 °C. However, the XRD patterns of samples synthesized at temperatures above 100 °C could be indexed to SnSe2 phase (JCPDS 01-089-2939) without any impure phase. Furthermore, the intensity of peaks becomes stronger and the corresponding full-width at half-maximum (FWHM) decreases with increasing reaction temperature. This result suggests that the elevated temperature activates the reaction and then increases the grain size of SnSe2 nanostructures. Similarly, it is unable to obtain SnSe nanostructures at temperature below 110 °C as reflected by Fig. 1(b). Fortunately, the XRD patterns of powders obtained at the temperature above 110 °C could be ascribed to SnSe crystal structure (JCPDS 00-048-1224). No obvious peaks of impurities can be observed when the reaction temperature exceeds 140 °C, demonstrating the successful synthesis of SnSe nanostructures via this wet-chemistry method.

Fig. 1. (color online) XRD patterns of (a) SnSe2 and (b) SnSe powders synthesized via a wet-chemistry route at different reaction temperatures.

The morphologies of the SnSe2 and SnSe nanostructures were observed by SEM and TEM. As given in Fig. 2, the characterization of SEM shows that the SnSe2 nanostructures obtained at 140 °C have the shape of nano-petal (Fig. 2(a)). On the other hand, the morphology of the sample synthesized at 182 °C looks more like potato chips with nanosize (Fig. 2(b)). Obviously, the increasing of reaction temperature increases the size of nanostructures, which is in accordance with the XRD results. In detail, the width of these two-dimensional nanostructures is expanded from nanometer (140 °C) to micrometer (182 °C). Though SnSe also has layered crystal structure, the SnSe nanostructures display agglomerated nanoparticles. Meanwhile, the sample obtained at 182 °C (Fig. 2(d)) presents larger agglomerations than that of the one prepared at 140 °C (Fig. 2(c)). Figure 3 shows the typical TEM and HRTEM images. In agreement with the observation of SEM, the SnSe2 sample synthesized at 182 °C (Fig. 3(a)) reveals nanosheets with the thickness of several nanometers to about 50 nm and the width of several hundred nanometers to about . In contrast, nanoparticles with the size of several nanometers to about 25 nm are absorbed on the surface of lumps in the SnSe sample obtained at 182 °C (Fig. 3(b)). The HRTEM characterization reconfirms the crystal structures of SnSe2 and SnSe by matching the crystal plane spacing as shown in the insets of Figs. 3(a) and 3(b), respectively.

Fig. 2. SEM images of SnSe2 and SnSe nanostructures synthesized at 140 °C [(a) and (c)] and 182 °C [(b) and (d)], respectively.
Fig. 3. (color online) TEM images of (a) SnSe2 and (b) SnSe nanostructures synthesized at 182 °C. The insets are their corresponding HRTEM images.

We prepared the polycrystalline SnSe via two methods as described in Section 2. The only difference between them is the choice of precursors, namely SnSe2 and SnSe nanostructures. The structure of the samples was investigated by XRD measurement. The collected patterns are given in Fig. 4 and demonstrate that the prepared SnSe2 keeps its original crystal structure until the temperature of heat-treating is elevated to 550 °C, where some peaks indexed to SnSe phase (JCPDS 00-014-0059) appear. With further increasing the temperature, the precursor of SnSe2 transforms into SnSe phase completely at 600 °C, without observable diffraction of impurities. Due to the higher vapor pressure of Se than that of Sn, it is anticipated that the sublimation of Se should be stronger than that of Sn, driving the formation of SnSe phase at high temperature ( ). When the temperature reaches 850 °C, a small amount of powders were collected in our experiments, which is not enough to fabricate bulk materials. Interestingly, the obtained SnSe phase indicates a strong anisotropic crystal growth, which is reflected by the fact that the diffraction intensity of (400) crystal face is particularly stronger than the others. The bc orientation degree for the (h00) crystal planes termed as F (h00) can be calculated with Lotgering method by the following equations[42] where P and P0 are the ratios of the integrated intensities of all (h00) crystal planes to those of all (hkl) planes for preferentially and randomly oriented samples, respectively. F = 0 and F = 1 mean completely disordered and ordered, respectively. In the present work, the P and P0 are calculated by using the XRD data and the standard card (JCPDS No. 00-014-0159). The result indicates that the values are 0.51, 0.76, and 0.89 for the XRD patterns of the samples with heat-treating temperature of 600 °C, 700 °C, and 800 °C, respectively. This improved anisotropic crystal growth along bc plane should be beneficial to improving the thermoelectric properties since the direction along b axis has the best thermoelectric performance.[17] Oppositely, the strongest diffraction peak of the samples prepared by using SnSe nanostructures as the precursor could be ascribed to (111) crystal face, which is typically shown in the inset in Fig. 4.

Fig. 4. (color online) XRD patterns of the polycrystalline obtained by treating the precursors of SnSe2 nanostructures in a tubular furnace at different temperatures. The inset is the XRD patterns of the reference, which prepared by using SnSe nanostructures as the precursor.

The phenomena of anisotropic crystal growth are also affirmed by the SEM observation, which was performed on the fracture surface parallel to the pressing direction of the SPS process. Figure 5(a)5(c) give the images of the sample 400 °C@SnSe2, 600 °C@SnSe2, and 800 °C@SnSe2, respectively. Apparently, all samples show a layered morphology. Meanwhile, the size of slice becomes larger and larger with increasing the temperature. This is reasonable because that both SnSe2 and SnSe have a layered structure. However, as shown in Fig. 5(d), the anisotropic crystal growth of the sample 600 °C@SnSe is weaker.

Fig. 5. (color online) SEM images of the fracture surface parallel to the pressing direction of SPS process: (a) 400 °C@SnSe2, (b) 600 °C@SnSe2, (c) 800 °C@SnSe2, and (d) 600 °C@SnSe.

Figure 6 displays the thermoelectric properties being measured perpendicular to the pressing direction of the SPS process, which present a better thermoelectric performance than those being measured the parallel to the pressing direction in our experiment. In Fig. 6(a), we find that the electrical resistivity increases gradually at temperature ranging from 50 °C to about 300 °C, and then decreases sharply at temperature ranging from 300 °C to about 550 °C for the SnSe bulk materials prepared by using SnSe2. Meanwhile, the increasing of the heat-treating temperature decreases the electrical resistivity from (600 °C) to (800 °C) at about 612 °C. This phenomenon may be mainly owing to the increased grain size as observed by SEM, which should improve the carrier mobility due to the decreased carrier scattering at grain boundaries. On the other hand, the sample 600 °C@SnSe has higher electrical resistivity. Generally, the electrical conductivity is determined by the carrier concentration and mobility according to the formula . Therefore, anisotropic crystal growth in polycrystalline SnSe along bc plane would favor the transport of carriers because the carrier mobility in SnSe crystals is higher along b and c axes than that of a axis. Based on the observation of SEM, which shows that the sample 600 °C@SnSe has weaker anisotropic crystal growth than the others, it is easy to anticipate that the carrier mobility in sample 600 °C@SnSe would be lower. Besides, it is reasonable to conclude that the sample T@SnSe2 contains more Sn vacancy than T@SnSe, since the former (1.12/1) has a higher / value than that of the latter one (1.03/1) as shown in Table 1. Usually, Sn vacancy could provide holes,[34] and then both the increased carrier mobility and concentration in sample T@SnSe2 would decrease the electrical resistivity. The evaluated Hall concentration and mobility at room temperature as shown in Table 2 also agree well with the above conclusion as follows. The sample 600 °C@SnSe2 ( , ) owns higher carrier concentration and mobility than the sample 600 °C@SnSe ( , ), suggesting that the precursor of SnSe2 nanostructures is more beneficial for the generation of Sn vacancy and anisotropic crystal growth. At the same time, the elevated heat-treating temperature may also enhance the formation of Sn vacancies and anisotropic crystal growth since the carrier concentration and mobility are improved to and for the sample 800 °C@SnSe2, respectively. The Seebeck coefficient measurement given in Fig. 6(b) shows that SnSe2 bulk materials have negative values, suggesting that they are n-type semiconductors, while polycrystalline SnSe bulk materials presents positive values, reflecting their p-type properties. This is consistent with the results in earlier reports.[34,40] Besides, the Seebeck coefficient depending on the heat-treating temperature varies limitedly. For example, the Seebeck coefficient of the sample 800 °C@SnSe2 is while that of the sample 700 °C@SnSe2 is at 612 °C. In line with the trend of electrical resistivity, the Seebeck coefficient of the sample 600 °C@SnSe is reasonably higher. However, because of its higher electrical resistivity, the sample 600 °C@SnSe shows very low power factor as revealed in Fig. 6(c), only at 537 °C. In the samples T@SnSe2, thanks to the remarkably improved electrical transport, the power factor is enhanced from to at 612 °C, which is comparable with the single crystal.[17] The thermal diffusivity and thermal conductivity are shown in Figs. 6(d) and 6(e), respectively. From these figures, it is easy to conclude that the thermal transport in SnSe is more difficult than that in SnSe2. For instance, in comparison with the sample 800 °C@SnSe2 (1.32 W/mK), the thermal conductivity of the sample 400 °C@SnSe2 is almost twice at 51 °C ( ). As shown in the inset of Fig. 6(e), the lattice thermal conductivity of the sample 700 °C@SnSe2 locates only a little bit below that of the sample 800 °C@SnSe2 above 300 °C. Meanwhile, the thermal conductivity of the sample 800 °C@SnSe2 is also relatively low, only at 612 °C. As a result of the greatly improved power factor and relatively low thermal conductivity, a maximum zT value of 1.07 at 612 °C is obtained by the sample 800 °C@SnSe2. To the best of our knowledge, it is one of the best results of SnSe polycrystalline samples without foreign-atomic doping.

Fig. 6. (color online) Thermoelectric properties of the polycrystalline along the orientation perpendicular to the pressing direction of SPS process: (a) electrical resistivity, (b) Seebeck coefficient, (c) power factor, (d) thermal diffusivity, (e) thermal conductivity ( , and (f) zT. The inset in panel (e) is the lattice thermal conductivity ( , which is calculated based on . The carrier thermal conductivity ( is estimated based on the Wiedemann–Franz law ( ). The Lorenz number is derived from Fermi energy with a simple one-band model and only acoustic phonon scattering considered.
Table 1.

The chemical composition of the polycrystalline samples obtained by XRF measurement.

.
Table 2.

The carrier concentration and Hall mobility of the polycrystalline samples.

.
4. Conclusion

In this study, by using SnSe2 nanostructures as the precursor, polycrystalline SnSe was successfully prepared by a combined heat-treating process in a reducing atmosphere and following spark plasma sintering procedure It shows better thermoelectric properties than that of the reference sample 600 °C@SnSe. Owing to the reduction of electrical resistivity, the power factor is enhanced from to at 612 °C. XRD, SEM, XRF, and Hall effect characterizations suggest that anisotropic crystal growth and Sn vacancy may be responsible for the enhanced electrical transport. In addition, the thermal conductivity has a weak relationship with the heat-treating temperature As a result, a high zT value of 1.07 at 612 °C is achieved, which is one of the best results of SnSe polycrystalline samples without foreign-atomic doping. This study provides a new method to synthesize polycrystalline SnSe and pave a way to improve the thermoelectric properties of polycrystalline bulk materials with similar layered structure.

Reference
[1] Liu W Jie Q Kim S H Ren Z 2015 Acta Mater. 87 357
[2] Alam H Ramakrishna S 2013 Nano Energy 2 190
[3] Pei Y Gibbs Z M Gloskovskii A Balke B Zeier W G Synder G J 2014 Adv. Energy Mater. 4 1400486
[4] Zhao L D Tan G Hao S He J Pei Y Chi H Wang H Gong S Xu H Dravid V P Uher C Snyder G J Wolverton C Kanatzidis M G 2016 Science 351 141
[5] Pei Y Shi X LaLonde A Wang H Chen L Snyder G J 2011 Nature 473 66
[6] Peng K Zhang B Wu H Cao X Li A Yang D Lu X Wang G Han X Uher C Zhou X 2017 Mater. Today 23 23
[7] Heremans J P Jovovic V Toberer E S Saramat A Kurosaki K Charoenphakdee A Yamanaka S Snyder G J 2008 Science 321 554
[8] Tan G Zhao L D Kanatzidis M G 2016 Chem. Rev. 116 12123
[9] Zhang X Liu J Li Y Su W B Li J C Zhu Y H Li M K Wang C M Wang C L 2015 Chin. Phys. Lett. 32 037201
[10] Wu D Wu L He D Zhao L D Li W Wu M Jin M Xu J Jiang J Huang L Zhu Y Kanatzidis M G He J 2017 Nano Energy 35 321
[11] Cao B L Jian J K Ge B H Li S M Wang H Liu J Zhao H Z 2017 Chin. Phys. 26 017202
[12] Minnich A J Dresselhaus M S Ren Z F Chen G 2009 Energy Environ. Sci. 2 466
[13] Wu Z H Xie H Q Zhai Y B Gan L H Liu J 2015 Chin. Phys. 24 034402
[14] Zhao L D Dravid V P Kanatzidis M G 2014 Energy Environ. Sci. 7 251
[15] Biswas K He J Blum I D Wu C I Hogan T P Seidman D N Dravid V P Kanatzidis M G 2012 Nature 489 414
[16] Kleinke H 2010 Chem. Mater 22 604
[17] Zhao L D Lo S H Zhang Y Sun H Tan G Uher C Wolverton C Dravid V P Kanatzidis M G 2014 Nature 508 373
[18] Li C W Hong J May A F Bansal D Chi S Hong T Ehlers G Delaire O 2015 Nat. Phys. 11 1063
[19] Peng K Lu X Zhan H Hui S Tang X Wang G Dai J Uher C Wang G Zhou X 2016 Energy Environ. Sci. 9 454
[20] Fu J Su X Xie H Yan Y Liu W You Y Cheng X Uher C Tang X 2018 Nano Energy 44 53
[21] Ge Z H Song D Chong X Zheng F Jin L Qian X Zheng L Dunin-Borkowski R E Qin P Feng J Zhao L D 2017 J. Am. Chem. Soc. 139 9714
[22] Chere E K Zhang Q Dahal K Cao F Mao J Ren Z 2016 J. Mater. Chem. 4 1848
[23] Zhang Q Chere E K Sun J Cao F Dahal K Chen S Chen G Ren Z 2015 Adv. Energy Mater 5 1500360
[24] Chang C Tan Q Pei Y Xiao Y Zhang X Chen Y X Zheng L Gong S Li J E He J Zhao L D 2016 RSC Adv. 6 98216
[25] Wang X Xu J Liu G Fu Y Liu Z Tan X Shao H Jiang H Tan T Jiang J 2016 Appl. Phys. Lett. 108 083902
[26] Ibrahim D Vaney J B Sassi S Candolfi C Ohorodniichuk V Levinsky P Semprimoschnig C Dauscher A Lenoir B 2017 Appl. Phys. Lett. 110 032103
[27] Zhao L D Chang C Tan G Kanatzidis M G 2016 Energy Environ. Sci. 9 3044
[28] Zhang B Peng K Sha X Li A Zhou X Chen Y Deng Q Yang D Ma E Han X 2017 Microsc. Microanal. 23 173
[29] Wei P C Bhattacharya S He J Neeleshwar S Podila R Chen Y Y Rao A M 2016 Nature 539 E1
[30] Wei T R Tan G J Zhang X Wu C F Li J F Dravid V P Snyder G J Kanatzidis M G 2016 J. Am. Chem. Soc. 138 8875
[31] Chen Y X Ge Z H Yin M J Feng D Huang X Q Zhao W Y He J Q 2016 Adv. Funct. Mater 26 6836
[32] Chattopadhyay T Pannetier J Vonschnering H G 1986 J. Phys. Chem. Solids 47 879
[33] Asfandiyar A Li Z L Sun F H Pan Y Wu C F Farooq M U Tang H C Li F Li B Li J F Wei T R 2017 Sci. Rep. 7 43262
[34] Wang X Xu J Liu G Q Tan X Li D Shao H Tan T Jiang J 2017 NPG Asia. Mater 9 e426
[35] Popuri S R Pollet M Decourt R Morrison F D Bennett N S Bos J W G 2016 J. Mater. Chem. 4 1685
[36] Wei W Chang C Yang T Liu J Tang H Zhang J Li Y Xu F Zhang Z Li J F Tang G 2018 J. Am. Chem. Soc. 140 499
[37] Dewandre A Hellman O Bhattacharya S Romero A H Madsen G K H Verstraete M J 2016 Phys. Rev. Lett. 117 276601
[38] Liu C Miao L Hu D Huang R Fisher C A J Tanemura S Gu H 2013 Phys. Rev. 88 205201
[39] Saha S Banik A Biswas K 2016 Chem. Eur. J. 22 15634
[40] Luo Y Zheng Y Luo Z Hao S Du C Liang Q Li Z Khor K A Hipalgaonkar K Xu J Yan Q Wolverton C Kanatzidis M G 2017 Adv. Energy Mater. 23 1702167
[41] Xu P Fu T Xin J Liu Y Ying P Zhao X Pan H Zhu T 2017 Sci. Bull. 62 1663
[42] Li F Zheng Z Li Y Wang W Li J F Li B Zhong A Luo J Fan P 2017 J. Mater Sci. 52 1