Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

pubs.acs.

org/journal/ascecg Research Article

High-Entropy Perovskites: An Emergent Class of Oxide


Thermoelectrics with Ultralow Thermal Conductivity
Ritwik Banerjee,# Sabitabrata Chatterjee,# Mani Ranjan, Tathagata Bhattacharya, Soham Mukherjee,
Subhra Sourav Jana, Akansha Dwivedi, and Tanmoy Maiti*
Cite This: ACS Sustainable Chem. Eng. 2020, 8, 17022−17032 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV PARIS-SUD on March 2, 2021 at 10:23:34 (UTC).

ABSTRACT: Although SrTiO3-based perovskites showed a lot of promise as n-type thermoelectric (TE) materials, they
demonstrated a low figure of merit value primarily because of their high lattice thermal conductivity (kl). Researchers found it
difficult to reduce kl, as a popular route like nanostructuring did not work well with these perovskites possessing low phonon mean
free paths. Here, we put forward a novel strategy of designing high-entropy perovskite (HEP) oxides having five transition metals in
the B site to induce more anharmonicity causing enhanced multiphonon scattering in order to decrease kl. Using detailed
thermodynamic calculations, we designed and synthesized a highly dense Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP ceramic. An ultralow
thermal conductivity of 0.7 W/mK at 1100 K was achieved in this n-type rare-earth-free HEP oxide TE material. The concept of
designing HEPs to achieve ultralow thermal conductivity potentially opens up a new avenue for enhancing TE performance of
environmentally benign bulk oxides for high-temperature TE power generation.
KEYWORDS: high entropy, perovskites, oxides, thermoelectric, low thermal conductivity

■ INTRODUCTION
High-entropy alloys (HEAs),1−3 whose essence lies in the
effect. However, these reports on HEP bulk oxides hardly
explored their functional properties, except the one on magnetic
mixing of five or more elements in equal proportion giving rise to behavior.16 Herein, we posited the concept of HEPs for
a single-phase solid solution, resulting in favorable Gibbs free designing novel thermoelectric (TE) materials. TE materials
energy by overcoming the usual phase separation caused by the
are considered for clean energy generation using the waste heat
high enthalpy effect,4 has demonstrated exciting properties in
recent times.5 HEAs showed promise for various applications in produced by manufacturing industries, furnaces, combustion
transportation and automobile industries,6 coatings,7 and so on, engines, various electronic appliances, and so on. Energy
due to their better performance than that of constituent conversion efficiency of a TE power generator is typically
elements.8 In 2015, Rost et al.9 developed high-entropy oxides dictated by a single dimensionless quantity called figure of merit
by incorporating five cations equiatomically in an oxide structure
(ZT):
to convert the multiphase structure into a single phase, utilizing
the similar concept of HEAs. Later, few more reports on the
electrical and magnetic properties of high-entropy oxides were Received: May 23, 2020
made in the last 3 years.10−13 Over the past one year, high- Revised: October 16, 2020
entropy perovskites (HEPs) or HEP oxides14−16 have been Published: November 5, 2020
developed by researchers, where the A site or B site of the ABO3
perovskite structure were populated with five cations, most of
the time mixed equiatomically to maximize the high-entropy

© 2020 American Chemical Society https://dx.doi.org/10.1021/acssuschemeng.0c03849


17022 ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 1. (a) Schematic representation of the crystal structure of Sr(TiFeNiCrNb)O3. (b) Estimated configurational entropy as a function of mol % of
the n-th component in the solid solution of n-component systems. The black dots indicate the compositions showing the maximum configurational
entropy.

S2σ S2σ conductivity as well as the Seebeck coefficient. In the current


ZT = T= T investigation, we synthesized a novel HEP oxide, Sr-
κ ke + kl (1)
(Fe0.2Ti0.2Mo0.2Nb0.2Cr0.2)O3, using the solid-state reaction
where S is the Seebeck coefficient, σ is the electrical conductivity, method and evaluated its potential for high-temperature TE
and k is thermal conductivity. k is written as a combination of applications in the temperature range from 300 to 1100 K. The
electronic (ke) and lattice thermal conductivity (kl). Oxides were HEP was designed by calculating Goldschmidt’s tolerance factor
contemplated as promising TE materials17,18 for high-temper- and configurational entropy. An ultralow lattice thermal
ature power generation due to their low processing cost, high conductivity of 0.7 W/mK at 1050 K was obtained in the
temperature durability, and environmental friendliness com- Sr(Fe0.2Ti0.2Mo0.2Nb0.2Cr0.2)O3 HEP by inducing enhanced
pared with intermetallics and chalcogenides. Although there are electron−phonon and Umklapp scattering, determined by the
some cobaltite-based p-type oxides with good ZT values,19−21 Debye−Callaway model. To the best of our knowledge, this is
scientists found it hard to design high-performance n-type TE the first report on TE properties of HEPs.


oxide materials. Nevertheless, for the efficient energy con-
version, a TE generator22 requires both n-type and p-type TE DESIGN AND THERMODYNAMIC CALCULATIONS
materials, where they are connected thermally in parallel and OF HEPS
electrically in series. Among various n-type oxide materials,
SrTiO3-based perovskites doped with La and Nb have shown a Over the years, scientists have relied on Goldschmidt’s tolerance
lot of promise in recent years.23−26 However, these perovskites factor to design novel ABO3 perovskite structures. In order to
suffer from high thermal conductivity (>10 W/mK),27−29 check the stability of the Sr(Fe0.2Ti0.2Mo0.2Nb0.2Cr0.2)O3 HEP
resulting in poor ZT values compared with those of the state- structure as shown schematically in Figure 1a in the present
of-the-art bulk TE materials like tellurides, which exhibit k less work, we further considered the expression of tolerance factor
than 1 W/mK.30,31 Because lattice thermal conductivity (kl) (t) as follows:
accounts for the lion’s share of total thermal conductivity of
rA + rO rSr + rO
these materials, it is of utmost concern to strategize for inducing t= = rFe + rTi + rMo + rNb + rCr
enhanced phonon scattering in order to achieve higher ZT
values in oxide perovskite systems. The nanostructuring
2 (rB + rO) 2 ( 5 )
+ rO

approach was proven to be the effective way for reducing kl in = 0.906 (2)
intermetallic- and chalcogenide-based TE materials by increas-
ing grain boundary scattering.32 However, the same strategy where rA, rB, rSr, rFe, rTi, rMo, rNb, rCr, andrO are the radii of cations
cannot be implemented in SrTiO3-based perovskite systems A, B, Sr, Fe, Ti, Mo, Nb, Cr, and anion O, respectively. A cubic
unless a microstructure with nm-scale grain size is achieved phase is likely to be stable when 0.9 ≤ t ≤ 1, while an
because the phonon mean free path of these perovskites typically orthorhombic or rhombohedral phase may form when t<0.9,
lies around <2 nm.33 Hence, the attention of the scientific and a tetrahedral or hexagonal phase may form when t>1. From
community has been diverted34,35 toward evolving a new the calculated tolerance factor, it is apparent that the stable cubic
strategy to design novel oxide TE materials with intrinsically HEP crystal structure could be formed in the present work.
ultralow lattice thermal conductivity, similar to what is generally Thermodynamics calculations were performed to understand
found in the state-of-the-art TE materials like tellurides, in order the formation mechanism of complex HEPs keeping in mind
to achieve high ZT values in oxide perovskites. Hess’ law at 1500 K. In the present work, precursors such as
Herein, we propose that designing HEPs by colonizing the B SrCO3, Fe2O3, MoO3, TiO2, Nb2O5, and Cr2O3 were used to
site of SrTiO3 with five transition metals equiatomically could obtain the single-phase solid solution Sr-
potentially induce multiphonon scattering resulting in lower kl. (Fe0.2Ti0.2Mo0.2Nb0.2Cr0.2)O3 at 1500 K.
Also, being a multicationic oxide, the HEP structure could be Thermal enthalpy (H) and thermal entropy (S) of all the
tailored in such a way that it could exhibit good electrical oxides can be calculated as follows:
17023 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 2. (a) Experimental (open circles), calculated (continuous line), difference profile (continuous line), and Bragg position (vertical bars)
obtained after Rietveld refinement of XRD data. (b) FE-SEM image of the HEP ceramics. (c) EDXS of Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 ceramics
showing elemental mapping.

H = Ht0 − H298
0 Thermodynamic Gibbs free energy change (ΔGThermo) has
been calculated as −2.35KJ/mol, as depicted in the Supporting
= At + (Bt 2)/2 + (Ct 3)/3 + (Dt 4)/4 − E /t + F Information.
− Hf ΔGT is the total free energy change of the reaction,
(3)
summation of thermodynamic Gibbs free energy change
(ΔGThermo) and free energy change due to mixing (ΔGmix).
S = St0 = Alnt + Bt + Ct 2/2 + Dt 3/3 − E /(2t 2) + G
(4) ΔGmix = ΔHmix − T ΔSmix (7)
where A, B, C, D, andE are the substance-dependent constants In our system of HEP oxides, the A site (0,0,0) is always
and “t” is the temperature(K)/1000. occupied by Sr2+ ions and O2− ions always occupy the (1/2,1/
To calculate enthalpy of formation (Hf) for AaBbOy oxides, the 2,0) site. Hence, it is assumed that the mixing happens only in
natural extension rule was employed on Pauling’s proposed the B site (1/2,1/2,1/2). The B site has been populated by Ti,
formula36 slightly modifying the one proposed by Aronson.37 Fe, Mo, Nb, and Cr, and mixing enthalpy was calculated only for
B-site cations. The calculation of ΔHmix40,41 for HEAs is usually
Hf = −96.5 × y[xA(χA − χO )2 + x B(χB − χO )2 ] + 108.8y done using the equation originally proposed by Miedema.42 The
(5) detailed calculations for enthalpy of mixing and entropy of
mixing are provided in the Supporting Information. Free energy
where χ represents the electronegativity taken from Pauling
change due to mixing (ΔGmix) has been calculated as ΔGmix =
electronegativity series; x, the mole fraction; and y, the number
29.128KJ/mol.
of oxygen atoms (O) in the oxide molecule under consideration.
Finally, ΔGTotal can be calculated as follows:
Consequently, the natural extension for mixed oxide Sr-
(Ti0.2Mo0.2Fe0.2Nb0.2Cr0.2)O3 takes the following form: ΔGTotal = ΔGThermo + ΔGmix = −2.35 − 29.128
Hf = −96.5 × y[xSr(χSr − χO )2 + x Fe(χFe − χO )2 = −31.488 KJ/mol (8)
2 2
+ xMo(χSr − χO ) + x Nb(χSr − χO ) + xCr The abovementioned calculations have been further demon-
strated in Figure 1b, where it is shown that for any component,
(χSr − χO )2 + x Ti(χSr − χO )2 ] + 108.8y an equiproportional system gives the maximum entropy and
adding another element to that n-component system increases
= −987.93KJ/mol (6)
the entropy starting from the maximum entropy of the n-th
component system.


Putting back the value of Hf and using the boundary
conditions H = 0 at a temperature of 298 K and S = 0 at 0 K,
two integration constants such as F and G were calculated using EXPERIMENTAL PROCEDURE
eqs 3 and 4. We used the values of A, B, C, D, and E and the Dense samples of Sr(Fe0.2Ti0.2Mo0.2Nb0.2Cr0.2)O3 HEP were synthe-
formation enthalpy (Hf) for precursor oxides from earlier sized by the conventional solid-state reaction method. To obtain the
reports.38,39 By fitting the CP data of the HEP sample shown in stoichiometric composition of the HEP, powders of SrCO3 (≥99.9%;
Figure S1a and putting the boundary conditions, the values of A, Sigma-Aldrich), Fe2O3 (≥99.995%; Sigma-Aldrich), MoO3(≥99%;
Sigma-Aldrich), TiO2 (≥99.5%; Sigma-Aldrich), Nb2O5(≥99%;
B, C, D, E, F, and G were calculated for HEPs as presented in Sigma-Aldrich), and Cr2O3(≥99.5%; Sigma-Aldrich) were mixed in
Table S1 in the Supporting Information. Furthermore, we appropriate ratios in a ball-mill at 350 rpm for 24 h using ethanol as the
calculated the change in enthalpy (ΔH) and entropy (ΔS) by milling medium. The milled and dried powder was then calcined at
plugging all the values of constants in eqs 3 and 4, which is 1500 K for 10 h in a H2 atmosphere. The agglomerated powder
shown in the Supporting Information. obtained after the calcination process was milled again to breakdown

17024 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

the agglomerates in a planetary micro mill (PULVERISETTE 7 To quantify the composition of the equimolar HEP, we further
premium line; Fritsch, Rhineland-Palatinate, Germany) for 120 min at carried out the XRF study on the as-synthesized ceramic
600 rpm using zirconia grinding balls and ethanol as milling media. samples. XRF peaks corresponding to all the elements present in
After adding binders, the milled powder was uniaxially pressed into our HEP sample are shown in Figure S2 in the Supporting
pellets and sintered at 1550 K for 10 h in a H2 atmosphere. The X-ray
diffraction (XRD) patterns of HEP powders were measured in the
Information. Table 2 shows weight percentages of the different
range of 2θ = 10−100° using a PANalytical X’Pert diffractometer to
determine the phase purity of calcined and sintered powders. The Table 2. Chemical Composition of the HEP Sample Obtained
microstructure of the fracture surface was observed by field-emission from XRF Analysis
scanning electron microscopy (FE-SEM, Nova NanoSEM 450 N2),
and energy-dispersive X-ray spectroscopy (EDXS) analysis was done average chemical composition in wt %
using a WSEM (EV050; Carl Zeiss NTS GmbH, Germany) machine. element expected observed by XRF
X-ray fluorescence (XRF) analysis of the HEP sample was carried out to Sr 55.97 55.8 ± 0.5
estimate its chemical composition using a Rigaku WD-XRF machine (4
Ti 6.11 5.8 ± 0.3
kW, 60 kV, 150 mA). X-ray photoelectron spectroscopy (XPS) was
carried out using PHI 5000 VersaProbe II, FEI Inc. TE properties, such Cr 6.64 6.5 ± 0.3
as electrical conductivity (σ) and Seebeck coefficient (S), were Fe 7.13 7.0 ± 0.3
simultaneously measured in the temperature range from 300 to 1100 Nb 11.87 11.8 ± 0.4
K at an interval of 50 K using a ZEM-3M10R apparatus (ULVAC- Mo 12.26 12.8 ± 0.5
RIKO Inc.). Subsequently, the TE power factor S2σ was calculated for
all the compositions in the temperature range from 300 to 1100 K.
Density (ρ) of the sample was measured by Archimedes’ principle using
a custom-made density measurement kit. Thermal diffusivity (D) and elements present in the HEP composition. It can be concluded
specific heat (Cp) were measured in the temperature range from 300 to from the XRF study that almost an equimolar ratio of the five
1100 K using LFA-1000 (LINSEIS) and HDSC PT 1600 (LINSEIS) transition metals were obtained from our Sr-
instruments, respectively, to calculate the thermal conductivity (κ =
ρDCp). Nanoindentation was done using a Nanoindenter with in situ
(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP sample, as expected.
scanning probe microscopy (TI 750; Hysitron Ltd.). XPS confirmed the presence of all the cations in the B site,


such as Ti, Fe, Mo, Nb, and Cr, possessing multiple oxidation
RESULTS AND DISCUSSION states. All the peaks were fitted using Lorentzian−Gaussian line
shapes after background correction as shown in Figure 3. The
Structural Analysis. The XRD pattern of the sintered calculated binding energies for all the transition metals present
Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP sample suggests the in HEPs are presented in Table 3. Figure 3a shows the existence
formation of a single-phase solid solution with a cubic perovskite of Ti in the form of ∼19% Ti2+, ∼58% Ti3+, and ∼23% Ti4+.
structure as no secondary phases or peak splitting was detected, Similarly, Fe was found to be present in the form of Fe2+ (∼72%)
shown in Figure 2a. Furthermore, we carried out Rietveld and Fe3+ (∼28%) as shown in Figure 3b. Mo was also found to
analysis of the XRD data using Fullprof software to determine possess a mixed valence state as Mo5+ (∼43%) and Mo6+
the crystal structure of the HEP sample. Figure 2a shows the (∼57%). XPS also revealed that ∼40% of Nb was present as
observed, calculated, and difference profiles obtained after Nb4+ and 60% of Nb as Nb5+. Interestingly, 17% Cr was detected
Rietveld refinement. This refinement analysis attributes to in the form of Cr6+ along with 83% Cr3+ in our HEP sample. All
excellent fitting (χ2 = 1.75) with a cubic perovskite structure the binding energies calculated for the HEP sample were found
possessing a Pm3̅m space group as shown in Figure 2a. The to be close to the reported values in the literature.44−54 The
lattice constant and other fitting parameters obtained from presence of higher oxidation states of Nb, Cr, and Mo occupying
Rietveld refinement are presented in Table 1. the B site of the perovskite structure, which has the formal
valency of +4, probably caused the reduced oxidation states of Ti
Table 1. Refined Structural Parameters for the HEP and Fe.
Composition Obtained after Rietveld Refinement of the XRD Thermoelectric Properties. The temperature-dependent
Data Using aPm3̅m Space Groupa Seebeck coefficient (S) of the HEP sample is plotted as shown in
atoms X Y Z B (Å2) Figure 4a. The Seebeck coefficient of the HEP was found to be
negative and increasing in magnitude in the temperature range
Sr 0.0 0.0 0.0 0.43
Ti/Nb/Fe /Cr/Mo 0.5 0.5 0.5 0.40
from 300 to 1100 K. It suggests the n-type TE behavior, that is,
O1 0.5 0.0 0.5 0.38
electrons being the majority charge carriers in a HEP system,
O2 0.5 0.5 0.0 1.74
which can be further analyzed by considering the defect
chemistry involved in this high-entropy ceramics. Equimolar
a
a = b = c = 3.94 Å; χ2 = 1.75; RF = 2.08; RP = 14.8; RWP = 8.06: RB = site occupancy by multivalent transition metals such as Ti, Fe,
1.09.
Mo, Cr, and Nb in the B site of ABO3 perovskites, which
formally occupies a charge valence of +4, gives rise to the
Microstructural studies of the HEP sample carried out by FE- formation of point defects explained by the following possible
SEM ascertained the formation of the dense ceramic pellet with defect reactions 9−13 expressed in the form of Kroger−Vink
low porosity as shown in Figure 2b. Micron-sized fused grains notation:55
were detected in the microstructure of HEP. The average grain BO2
MO2 ⎯⎯⎯→ M×B + 2OO
×
; For M = Ti4 +, Nb4 + (9)
size was estimated to be 0.85 ± 0.3 μm with the help of ImageJ
software.43 EDXS was done to check the compositional 2BO2 1
homogeneity of HEPs, and all the elements (Sr, Ti, Fe, Mo, M 2O5 ⎯⎯⎯⎯→ 2 M·B + 4 OO
×
+ O2 ↑ +2e′;For M
2
Nb, Cr, and O) were found to be uniformly distributed
throughout the area of the microstructure as shown in Figure 2c. = Nb5 + , Mo5 + (10)

17025 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 3. XPS spectra of (a) Ti 2p, (b) Fe 2p, (c) Nb 3d, (d) Mo 3d, and (e) Cr 2p peaks for the Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP.

Table 3. Measured Binding Energies of all the Elements 2BO2 ×


M 2O3 ⎯⎯⎯⎯→ 2M′B + 3 OO + V ··O; For M = Fe3 +, Cr 3 +
Present in the HEP Sample
(12)
element state binding energy(eV) area (%)
BO2
Mo Mo6+ 3d5/2 232.13 42.1 ×
MO ⎯⎯⎯→ M″O + OO + V ··O; For M = Fe 2 + (13)
Mo6+ 3d3/2 235.38 14.8
Mo5+ 3d5/2 231.36 25.3 It can be seen from the defect reactions 12 and 13 that lower
Mo5+ 3d3/2 234.54 17.8 oxidation states (<4) of transition metals auger the formation of
Fe Fe2+ 2p3/2 710.46 32.8 oxygen vacancies. On the other hand, higher oxidation states
Fe2+ 2p1/2 724.16 31.5 (>4) of Nb, Mo, and Cr occupying the B site generate electrons
Fe3+ 2p3/2 712.90 23.2 in the HEP samples as shown by eqs 10 and 11. Because we
Fe3+ 2p1/2 728.54 2.7 calcined as well as sintered our HEP samples in a reducing
satellite 2p3/2 717.94 1.6 atmosphere (H2), these defect reactions are expected to be
satellite 2p1/2 731.34 8.2 favored. Occurrence of a negative Seebeck coefficient suggest s
Cr Cr3+ 2p3/2 574.12 50.1 that the defect reactions 12 and 13 dominated in HEP ceramics.
Cr3+ 2p1/2 584.05 33.1 Furthermore, a monotonic increase in “S” with temperature
Cr6+ 2p3/2 577.01 15 indicates increased scattering of electrons with temperature
Cr6+ 2p1/2 587.64 1.8 implying a metallic or degenerate semiconductor-like behavior
Nb Nb4+ 3d5/2 206.12 21.9 of HEP samples.
Nb4+ 3d3/2 208.83 17.3 However, HEPs do not exhibit similar kind of metallic
Nb5+ 3d5/2 206.68 39.4 behavior in the temperature-dependent electrical conductivity
Nb5+ 3d3/2 209.36 21.4 graph as shown in Figure 4b. Indeed, electrical conductivity (σ)
Ti Ti2+ 2p3/2 457.2 11.3 was found to be continuously increasing from ∼500 S/m at 300
Ti2+ 2p1/2 462.06 7.7 K to ∼7.683 × 103 S/m at 1100 K, demonstrating a
Ti3+ 2p3/2 457.94 35.6 semiconductor-like (dσ/dT > 0) behavior. This can be
Ti3+ 2p1/2 463.72 22.7 explained by Anderson’s localization.56−58 Existence of multi-
Ti4+ 2p3/2 458.55 17.2 valent transition metals in HEPs creates lot of defect centers as
Ti4+ 2p1/2 465.59 5.4 explained by eqs 9−13 above, due to which large variations of
the local electrical field and lattice strain are expected to be
BO2 1
MO3 ⎯⎯⎯→ M··B + 2OO
×
+ O2 ↑ +2e′;For M present in the system. As a result, itinerant electrons can become
2 localized, which is called Anderson’s localization. Such kind of
= Mo6 +, Cr 6 + (11) behavior has been reported in the literature for perovskites57−59
and double perovskites.60−62 It is believed that the presence of
17026 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 4. Temperature-dependent (a) Seebeck coefficient (S), (b) electrical conductivity (σ), (c) power factor (S2σ), and (d) plot of ln(σT) versus 1/
KBT demonstrating the small polaron hopping (SPH) conduction mechanism in HEPs.

kB ij 2 − c yz
lnjj zz
five different transition metals possessing multiple oxidation
e k c {
S=
states caused the Anderson localization of charge carriers in (15)
HEPs. Hence, HEPs were not found to follow the degenerate
Fermi gas model in the temperature-dependent electrical where S is the Seebeck Coefficient at room temperature, kB is
conductivity behavior. As a result, we could achieve a rare feat Boltzmann’s constant, e is the electronic charge, and c is the
of simultaneous increases in both the Seebeck coefficient and fractional polaron concentration. For our HEP sample, the
fractional polaron concentration was calculated to be 0.71.
electrical conductivity with temperature, which potentially
Taguchi et al.65 calculated the number of carrier concen-
opens up a new way of designing oxide TE materials with an n
trations (n) using c = N in the original Heikes’ formula, which
enhanced power factor (S2σ). Figure 4c shows the continuous

ÄÅ ÉÑ
was further used in the recent works,66,67 given below:

kB ÅÅÅ ST* ÑÑÑ


increasing trend of TE power factor (S2σ) of the HEP sample
i
j y
z
S = ÅÅÅln β jj zz + ÑÑ
with temperature, where the highest value of 67.88 μW/mK2

e ÅÅÇ k c { KB ÑÑÑÖ
1 − c
was achieved at ∼1123 K.

Ä ÉÑ
Because certain activation energy barriers need to be (16)

A s ÅÅÅÅ ÑÑ
ÑÑ
overcome in order to delocalize the charge carriers trapped
Å
ÅÅ Ñ
V ÅÅÇ 1 + e(S × e)/ kB ÑÑÑÖ
around the defect sites, the SPH conduction model would be 1
appropriate for explaining the conduction mechanism in this n=
(17)

ÄÅ ÉÑ
HEP oxide, which is given by the following equation:63

σ0 ÅÅÅ −E H ÑÑÑ
where ST* is the vibrational entropy and is usually neglected, β is
σ = expÅÅÅ ÑÑ
ÅÅÇ KBT ÑÑÑÖ
the degeneracy factor taken as 1, and N is the number of
T (14) available sites per unit volume, further expressed as N = AS V . In
the present work, we used the following expression derived
where σ0 is a constant, EH is the activation energy of SPH, and KB similarly from the modified Heikes’ formula (eq 18) to calculate

Ä ÉÑ
A s ÅÅÅÅ
electron concentration:
ÑÑ
is Boltzmann’s constant. Figure 4d depicts the linear relationship

ÑÑ
ÅÅÅ ÑÑ
between ln(σT) and 1/KBT, indicating that the conduction
V ÅÅÇ 1 + e ÑÑÖ
2
mechanism of HEPs followed the SPH model in the entire range n= (S × e)/ k
of measurement temperature with an activation energy of (0.16
B
(18)
± 0.002) eV. When the charge transport is driven by SPH, the For cubic perovskites, the number of available sites per unit
effective charge carrier concentration can be determined using volume is taken as As= 1. Electron concentration in HEPs was
Heikes’ formula, which was further modified by Chaikin and estimated as 1.15 × 1022 per cm3. Furthermore, using σ = neμ,
Beni64 considering the spin interaction of electrons, elucidated mobility (μ) was calculated to be 0.0027 cm2/Vs. The effective
as follows: mass (m*) of electron was calculated from the Seebeck
17027 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 5. Variation of (a) total thermal conductivity (κ), (b) lattice thermal conductivity (κl), and (c) Lphonon with temperature for HEP composition.
(d) Comparative bar chart of Lphonon for SrTiO3-based perovskites in the temperature range of 300−1000 K.

coefficient expression for nondegenerate semiconductors (eqs The most interesting feature of this HEP system is its ultralow
19 and 20).68 lattice thermal conductivity of 0.7 W/mK exhibited at 1100 K.
To further gain insights into the lower lattice thermal
KB
S=− [ln( Nc n ) + A ] conductivity obtained in the HEP, the mean free path (Lphonon)
e (19) between two successive phonon collisions as a function of
temperature was estimated using following equation:70
Nc = 2( 2πm*KBT h2 )3/2 (20) 3 Kl
L Phonon =
Cv × vm (22)
where Nc is the density of states at the conduction band, h is
Planck’s constant, and A is the scattering parameter, ranging
where vm is the acoustic mean velocity in the medium and Cv is
between 0 and 4. Effective carrier mass (m*) for Sr-
the heat capacity at a constant volume. vm was calculated from
(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 was found to be 4.26m0, consid-
the elastic modulus measured by the nanoindentation method as
ering A = 3. The low value of electron mobility and high effective
described in the Supporting Information. The calculated Lphonon
mass further corroborate with Anderson’s localization of
electrons in the HEP system as discussed above. It provides is plotted as a function of temperature as shown in Figure 5c.
further opportunity to tune the composition of the perovskites Lphonon was found to be decreased from 0.8 nm at room
in order to delocalize these electrons which in turn can increase temperature to 0.5 nm at 1100 K. The values of Lphonon estimated
the electrical conductivity and power factor. for the HEP is further compared with that of the earlier reports
In order to calculate the total thermal conductivity (κ = for other SrTiO3-based oxide TEs33,34,71 as depicted by the
ρDCp), specific heat (Cp) and thermal diffusivity (D) of the HEP comparative bar chart shown in Figure 5d. In the literature, the
ceramic were measured as displayed in Figure S1a,b in the strategy of making nanocomposites, nanostructuring, and rare-
Supporting Information. Thermal conductivity (κ) of our HEP earth doping had been adopted to lower the mean free path in
ceramic decreased continuously with increasing temperature as perovskites in order to decrease their κl. However, we achieved
shown in Figure 5a. The primary reason behind such decreasing such low mean free path intrinsically in the single-phase solid
trend observed in κ is attributed to the monotonic decreasing solution of rare-earth-free HEP ceramics. This opens up the
behavior of dominant lattice thermal conductivity (κl) exhibited possibility of further reducing Lphonon, leading to lower Kl by
by this equimolar HEP oxide as shown in Figure 5b. κl was making composites with nanoscale grain size of HEPs.
calculated by subtracting the electronic thermal conductivity Reduction in lattice thermal conductivity is generally
(κe) from the total thermal conductivity (κ = κl + κe) of the governed by various kinds of phonon interactions, such as
material. The Wiedemann−Franz law (ke = LσT) was used to defect scattering, electron−phonon scattering, phonon−pho-
estimate the electronic thermal conductivity (κe) of the HEP non scattering. In order to understand the heat transport
oxide sample, shown in Figure S1c. The temperature-dependent mechanism in the HEP system, Callaway’s model was
Lorenz number was calculated using the following equation69 as considered. The Debye−Callaway model72 is given as

ÄÅ É kB ij kBT yz
Å −|S| ÑÑÑ
illustrated in Figure S1d:

jj zz
L = 1.5 + expÅÅÅÅ ÑÑ 2π 2vm jk ℏ z{
3 θD/ T
τ (y , T )y 4 e y
ÅÅÇ 116 ÑÑÑÖ (21)
kl = ∫
0
(e y − 1)2
dy
(23)

17028 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

where ℏ is the reduced Planck’s constant, kB is the Boltzmann Table 4. Fitting Parameters Obtained from the Analysis of
constant, τ is the relaxation time, vm is the average speed of Lattice Thermal Conductivity of
sound in that particular medium, θD is the Debye temperature Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP Using Callaway’s Model
calculated as 338 K, and y = ℏω/kBT. The calculation of Debye Described by Eqs 24 and 25 for Varying n Values
temperature is described in the Supporting Information.
n values A [s3] B [sK−1] C [s]
Because most of the measured values of thermal conductivity
−48
are in the range of temperature comparable with or greater than n=1 2.1 × 10 6.9 × 10−18 2.6 × 10−15
the Debye temperature (i.e., 338 K), we implied a more n=2 2.3 × 10−46 5.4 × 10−21 4.6 × 10−15
simplified form of the Callaway model derived by Slack73,74 as n=3 1.5 × 10−46 5.2 × 10−24 5.2 × 10−15
shown in eqs 24 and 25: n=4 3.5 × 10−46 5.0 × 10−27 5.6 × 10−15

kB ij kBT yz
n=5 3.8 × 10−46 4.9 × 10−30 5.9 × 10−15

jj zz
2π 2vm jk ℏ z{
3 θD/ T

kl = ∫
0
τ (y , T )y 2 d y
(24)
of higher-order phonon interaction processes in the HEP, which
might be arising due to increased anharmonicity occurred in the
vm HEP structure comprising five transition metals in the B site,
where τ −1 = + Aω4 + BωαT ne−θD/mT + Cω 2 compared with a simple perovskite system like SrTiO3.
L (25)
Noticeably, κ values obtained in the HEP were found to be
where A, B, and C are the constants, and relaxation time (τ) much lower than the best results available in the literature for
depends on grain boundary scattering ( vm ), defect or isotopes perovskites including SrTiO3-based TE materials34,78−81 as
L
scattering (Aω 4 ), phonon−phonon Umklapp scattering shown in Figure 7a. The obtained κ values are also found to be
lower than various perovskite ferroelectrics.82 This is phenom-
(BωαT ne−θD/mT ), and electron−phonon scattering (Cω2). At enal because the major drawback of perovskite-based TE
higher temperatures, the grain boundary scattering effects were materials is their high thermal conductivity. Our study shows
omitted, as suggested by Casimir.75 As suggested by Slack,76 for that designing rare-earth-free HEP-based oxides can be a
the Umklapp process, αis ≤ 2 and n is strongly temperature- potential way of fabricating oxide TE materials with ultralow
dependent. Furthermore, the exponential term in the Umklapp thermal conductivity. The presence of five multivalent cations at
process can be approximated to one for high temperature as the B site enhances the phonon scattering by introducing higher
suggested by Slack. The relaxation time for the three-phonon structural complexity in the HEP system, which in turn reduces
Umklapp process in the Callaway model was given by Bω2T3.76 the value of κl. Synergic enhancement of the Seebeck coefficient
However, in the modified Callaway model, the three-phonon and electrical conductivity along with a decrease in thermal
Umklapp process was shown as T-dependent rather than T3- conductivity with temperature is the ideal method to design TE
dependent. Slack further deduced the relaxation times for four- materials, and we achieved such kind of synergistic enhancement
phonon processes by considering T2 dependence.74 Recently, of TE performance in HEP ceramic systems. As a result, ZT of
Petersen et al.77 used four-phonon scattering processes by HEPs was found to be increased continuously with increasing
considering T2 dependence for a half-Heusler system. In the temperature as illustrated in Figure 7b. The highest ZT value
present work, α was fixed at 2, while n was varied from 1 to 5 achieved for HEPs is ∼0.065 at 1043 K, which requires much
until we get good fittings for temperature-dependent lattice improvement in order to make it viable for commercialization.
thermal conductivity data of HEPs using eqs 24 and 25 as shown Nevertheless, our study shows a new path of designing high-
in Figure 6 and Table 4. For n = 1 and 2, there was absolutely no performance perovskites, which are known to be amenable for
fitting observed. However, when fittings were done for n = 3, 4, compositional modification. Moreover, using HEPs can be the
and 5, the Callaway fit got better upon increasing the n value and potential way of decoupling “electron crystal” and “phonon
the best fit was obtained for n = 5. This implies the manifestation glass” in oxide TEs by careful selection of transition metals in
HEP systems.

■ CONCLUSIONS
In summary, we successfully presented the first comprehensive
study of the TE properties of HEP oxides. A novel perovskite
composition of five transition metals in the B site and a single
element in the A site, Sr.(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3, was
synthesized by the solid-state reaction route. A cubic crystal
structure possessing a Pm3̅m space group commensurate with
Goldschmidt’s tolerance factor was obtained. XRF analysis
revealed that the ratio of constituting elements is almost similar
to what was expected from the equimolar formula of HEPs. The
Seebeck coefficient signified HEPs to be an n-type material in
the entire temperature range from 323 to 1123 K. The
conductivity mechanism of the material was found to be driven
by the SPH model. Most importantly ultralow κ was achieved in
this HEP ceramic, which demonstrated synergistic enhancement
of the Seebeck coefficient and electrical conductivity. It is
Figure 6. Lattice thermal conductivity plotted as a function of anticipated that our HEP approach of designing oxide TEs will
temperature along with the Debye−Callaway model fitting for different lead to the discovery of novel TE materials with enhanced ZT
n values. values, which can be used for fabricating high-temperature
17029 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 7. (a) Thermal conductivity of the HEP compared with other SrTiO3-based composition reported in the literature and (b) variation of TE
figure of merit (ZT) for the Sr(Ti0.2Fe0.2Mo0.2Nb0.2Cr0.2)O3 HEP with temperature.

thermoelectric generator for electricity generation from waste Subhra Sourav Jana − Plasmonics and Perovskites Laboratory,
heat. Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh


208016, India
ASSOCIATED CONTENT Akansha Dwivedi − Department of Ceramic Engineering, Indian
Institute of Technology (BHU) Varanasi, Varanasi, Uttar
*
sı Supporting Information
Pradesh 221005, India
The Supporting Information is available free of charge at
Complete contact information is available at:
https://pubs.acs.org/doi/10.1021/acssuschemeng.0c03849.
https://pubs.acs.org/10.1021/acssuschemeng.0c03849
Calculation of Lorenz number, lattice thermal con-
ductivity (κl), and electrical thermal conductivity (κe); Author Contributions
#
the temperature-dependent plots of specific heat, thermal R.B. and S.C. equally contributed to this work.
diffusivity, and electronic thermal conductivity; XRF Notes
analysis; calculation of the Debye temperature taking into
The authors declare no competing financial interest.


consideration of bulk modulus, sheer modulus, and
Poisson’s ratio; and thermodynamic calculation: the
mixing enthalpy and entropy part of our HEP material ACKNOWLEDGMENTS
(PDF) This work is supported by the grant from the Science and


Engineering Research Board, DST (SERB-DST), India (Grant
No: IMP/20 l8/000955).


AUTHOR INFORMATION
Corresponding Author REFERENCES
Tanmoy Maiti − Plasmonics and Perovskites Laboratory, Indian (1) Jien-Wei, Y. Recent progress in high entropy alloys. Ann. Chim. Sci.
Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208016, Mater. 2006, 31, 633−648.
India; orcid.org/0000-0003-1581-7614; Email: tmaiti@ (2) Yeh, J. W.; Chen, S. K.; Lin, S. J.; Gan, J. Y.; Chin, T. S.; Shun, T.
iitk.ac.in T.; Tsau, C. H.; Chang, S. Y. Nanostructured high-entropy alloys with
multiple principal elements: novel alloy design concepts and outcomes.
Authors Adv. Eng. Mater. 2004, 6, 299−303.
Ritwik Banerjee − Plasmonics and Perovskites Laboratory, Indian (3) Cantor, B.; Chang, I. T. H.; Knight, P.; Vincent, A. J. B.
Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208016, Microstructural development in equiatomic multicomponent alloys.
India Mater. Sci. Eng., A Struct. Mater. 2004, 375-377, 213−218.
(4) Miracle, D. B.; Senkov, O. N. A critical review of high entropy
Sabitabrata Chatterjee − Plasmonics and Perovskites Laboratory, alloys and related concepts. Acta Mater. 2017, 122, 448−511.
Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh (5) Murty, B. S.; Yeh, J.-W.; Ranganathan, S.; Bhattacharjee, P. P.,
208016, India; Department of Ceramic Engineering, Indian High-Entropy Alloys. Elsevier: 2019, DOI: 10.1016/B978-0-12-
Institute of Technology (BHU) Varanasi, Varanasi, Uttar 816067-1.00010-2.
Pradesh 221005, India (6) Maulik, O.; Kumar, D.; Kumar, S.; Dewangan, S. K.; Kumar, V.
Mani Ranjan − Plasmonics and Perovskites Laboratory, Indian Structure and properties of lightweight high entropy alloys: a brief
Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208016, review. Mater. Res. Express 2018, 5, No. 052001.
India (7) Li, T.; Liu, Y.; Liu, B.; Guo, W.; Xu, L. Microstructure and Wear
Tathagata Bhattacharya − Plasmonics and Perovskites Behavior of FeCoCrNiMo0. 2 High Entropy Coatings Prepared by Air
Laboratory, Indian Institute of Technology Kanpur, Kanpur, Plasma Spray and the High Velocity Oxy-Fuel Spray Processes. Coatings
2017, 7, 151.
Uttar Pradesh 208016, India (8) Gorsse, S.; Nguyen, M. H.; Senkov, O. N.; Miracle, D. B. Database
Soham Mukherjee − Plasmonics and Perovskites Laboratory, on the mechanical properties of high entropy alloys and complex
Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh concentrated alloys. Data Brief 2018, 21, 2664−2678.
208016, India; Department of Ceramic Engineering, Indian (9) Rost, C. M.; Sachet, E.; Borman, T.; Moballegh, A.; Dickey, E. C.;
Institute of Technology (BHU) Varanasi, Varanasi, Uttar Hou, D.; Jones, J. L.; Curtarolo, S.; Maria, J.-P. Entropy-stabilized
Pradesh 221005, India oxides. Nat. Commun. 2015, 6, 8485.

17030 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

(10) Sarkar, A.; Velasco, L.; Wang, D.; Wang, Q.; Talasila, G.; de Biasi, (31) Pinisetty, D.; Devireddy, R. V. Thermal conductivity of
L.; Kübel, C.; Brezesinski, T.; Bhattacharya, S. S.; Hahn, H.; Breitung, B. semiconductor (bismuth−telluride)−semimetal (antimony) super-
High entropy oxides for reversible energy storage. Nat. Commun. 2018, lattice nanostructures. Acta Mater. 2010, 58, 570−576.
9, 3400. (32) Tan, G.; Zhao, L.-D.; Kanatzidis, M. G. Rationally designing
(11) Bérardan, D.; Franger, S.; Meena, A. K.; Dragoe, N. Room high-performance bulk thermoelectric materials. Chem. Rev. 2016, 116,
temperature lithium superionic conductivity in high entropy oxides. J. 12123−12149.
Mater. Chem. A 2016, 4, 9536−9541. (33) Wang, Y.; Fujinami, K.; Zhang, R.; Wan, C.; Wang, N.; Ba, Y.;
(12) Sarkar, A.; Wang, Q.; Schiele, A.; Chellali, M. R.; Bhattacharya, S. Koumoto, K. Interfacial thermal resistance and thermal conductivity in
S.; Wang, D.; Brezesinski, T.; Hahn, H.; Velasco, L.; Breitung, B. High- nanograined SrTiO3. Appl. Phys. Express 2010, 3, No. 031101.
Entropy Oxides: Fundamental Aspects and Electrochemical Properties. (34) Daniels, L. M.; Savvin, S. N.; Pitcher, M. J.; Dyer, M. S.; Claridge,
Adv. Mater. 2019, 31, 1806236. J. B.; Ling, S.; Slater, B.; Corà, F.; Alaria, J.; Rosseinsky, M. J. Phonon-
(13) Berardan, D.; Meena, A. K.; Franger, S.; Herrero, C.; Dragoe, N. glass electron-crystal behaviour by A site disorder in n-type thermo-
Controlled Jahn-Teller distortion in (MgCoNiCuZn) O-based high electric oxides. Energy Environ. Sci. 2017, 10, 1917−1922.
entropy oxides. J. Alloys Compd. 2017, 704, 693−700. (35) Daniels, L. M.; Ling, S.; Savvin, S. N.; Pitcher, M. J.; Dyer, M. S.;
(14) Jiang, S.; Hu, T.; Gild, J.; Zhou, N.; Nie, J.; Qin, M.; Harrington,
Claridge, J. B.; Slater, B.; Corà, F.; Alaria, J.; Rosseinsky, M. J. A and B
T.; Vecchio, K.; Luo, J. A new class of high-entropy perovskite oxides.
site doping of a phonon-glass perovskite oxide thermoelectric. J. Mater.
Scr. Mater. 2018, 142, 116−120.
(15) Sarkar, A.; Djenadic, R.; Wang, D.; Hein, C.; Kautenburger, R.; Chem. A 2018, 6, 15640−15652.
(36) Pauling, L., The Nature of the Chemical Bond. Cornell university
Clemens, O.; Hahn, H. Rare earth and transition metal based entropy
stabilised perovskite type oxides. J. Eur. Ceram. Soc. 2018, 38, 2318− press Ithaca, NY: 1960; 30, DOI: 10.1002/jps.3030300111.
2327. (37) Aronson, S. Estimation of the heat of formation of refractory
(16) Witte, R.; Sarkar, A.; Kruk, R.; Eggert, B.; Brand, R. A.; Wende, mixed oxides. J. Nucl. Mater. 1982, 107, 343−346.
H.; Hahn, H. High-entropy oxides: An emerging prospect for magnetic (38) Chase, M. W. NIST-JANAF thermochemical tables for oxygen
rare-earth transition metal perovskites. Phys. Rev. Mater. 2019, 3, fluorides. J. Phys. Chem. Ref. Data 1996, 25, 551−603.
No. 034406. (39) Jacob, K. T.; Shekhar, C.; Vinay, M.; Waseda, Y. Thermodynamic
(17) Walia, S.; Balendhran, S.; Nili, H.; Zhuiykov, S.; Rosengarten, G.; properties of niobium oxides. J. Chem. Eng. Data 2010, 55, 4854−4863.
Wang, Q. H.; Bhaskaran, M.; Sriram, S.; Strano, M. S.; Kalantar-zadeh, (40) Wang, X.-R.; He, P.; Lin, T.-S.; Wang, Z.-Q. Microstructure,
K. Transition metal oxides−Thermoelectric properties. Prog. Mater. Sci. thermodynamics and compressive properties of AlCrCuNixTi (x= 0, 1)
2013, 58, 1443−1489. high entropy alloys. Mater. Sci. Technol. 2015, 31, 1842−1849.
(18) Yin, Y.; Tudu, B.; Tiwari, A. Recent advances in oxide (41) Ye, Y. F.; Wang, Q.; Lu, J.; Liu, C. T.; Yang, Y. High-entropy
thermoelectric materials and modules. Vacuum 2017, 146, 356−374. alloy: challenges and prospects. Mater. Today 2016, 19, 349−362.
(19) Ohta, H.; Sugiura, K.; Koumoto, K. Recent progress in oxide (42) Zhang, R. F.; Zhang, S. H.; He, Z. J.; Jing, J.; Sheng, S. H.
thermoelectric materials: p-type Ca3Co4O9 and n-type SrTiO3−. Miedema Calculator: A thermodynamic platform for predicting
Inorg. Chem. 2008, 47, 8429−8436. formation enthalpies of alloys within framework of Miedema’s Theory.
(20) Rivas-Murias, B.; Vila-Fungueiriño, J. M.; Rivadulla, F. High Comput. Phys. Commun. 2016, 209, 58−69.
quality thin films of thermoelectric misfit cobalt oxides prepared by a (43) Schneider, C. A.; Rasband, W. S.; Eliceiri, K. W. NIH Image to
chemical solution method. Sci. Rep. 2015, 5, 11889. ImageJ: 25 years of image analysis. Nat. Methods 2012, 209, 671.
(21) Wu, N.; Linderoth, S.; Pryds, N.; Van Nong, N., Development and (44) Aronniemi, M.; Sainio, J.; Lahtinen, J. Chemical state
processing of p-type oxide thermoelectric materials. 2014. quantification of iron and chromium oxides using XPS: the effect of
(22) Hanson, C. M., Thermoelectric voltage generator. Google Patents : the background subtraction method. Surf. Sci. 2005, 578, 108−123.
1978. (45) Kadari, A.; Schemme, T.; Kadri, D.; Wollschläger, J. XPS and
(23) Li, J.-B.; Wang, J.; Li, J.-F.; Li, Y.; Yang, H.; Yu, H.-Y.; Ma, X.-B.; morphological properties of Cr2O3 thin films grown by thermal
Yaer, X.; Liu, L.; Miao, L. Broadening the temperature range for high evaporation method. Results Phys. 2017, 7, 3124−3129.
thermoelectric performance of bulk polycrystalline strontium titanate (46) Erdem, B.; Hunsicker, R. A.; Simmons, G. W.; Sudol, E. D.;
by controlling the electronic transport properties. J. Mater. Chem. C Dimonie, V. L.; El-Aasser, M. S. XPS and FTIR surface characterization
2018, 6, 7594−7603. of TiO2 particles used in polymer encapsulation. Langmuir 2001, 17,
(24) Wang, J.; Zhang, B.-Y.; Kang, H.-J.; Li, Y.; Yaer, X.; Li, J.-F.; Tan, 2664−2669.
Q.; Zhang, S.; Fan, G.-H.; Liu, C.-Y. Record high thermoelectric (47) Bharti, B.; Kumar, S.; Lee, H.-N.; Kumar, R. Formation of oxygen
performance in bulk SrTiO3 via nano-scale modulation doping. Nano vacancies and Ti 3+ state in TiO 2 thin film and enhanced optical
Energy 2017, 35, 387−395.
properties by air plasma treatment. Sci. Rep. 2016, 6, 32355.
(25) Roy, P.; Pal, V.; Maiti, T. Effect of Spark Plasma Sintering (SPS)
(48) Grosvenor, A. P.; Kobe, B. A.; Biesinger, M. C.; McIntyre, N. S.
on the thermoelectric properties of SrTiO3: 15 at% Nb. Ceram. Int.
Investigation of multiplet splitting of Fe 2p XPS spectra and bonding in
2017, 43, 12809−12813.
(26) Sun, J.; Singh, D. J. Thermoelectric properties of n-type SrTiO3. iron compounds. Surf. Interface Anal. 2004, 36, 1564−1574.
(49) Dupin, J.-C.; Gonbeau, D.; Vinatier, P.; Levasseur, A. Systematic
APL Mater. 2016, 4, 104803.
(27) Oh, D.-W.; Ravichandran, J.; Liang, C.-W.; Siemons, W.; Jalan, XPS studies of metal oxides, hydroxides and peroxides. Phys. Chem.
B.; Brooks, C. M.; Huijben, M.; Schlom, D. G.; Stemmer, S.; Martin, L. Chem. Phys. 2000, 2, 1319−1324.
W. Thermal conductivity as a metric for the crystalline quality of SrTiO (50) Grundner, M.; Halbritter, J. XPS and AES studies on oxide
3 epitaxial layers. Appl. Phys. Lett. 2011, 98, 221904. growth and oxide coatings on niobium. J. Appl. Phys. 1980, 51, 397−
(28) Okuda, T.; Nakanishi, K.; Miyasaka, S.; Tokura, Y. Large 405.
thermoelectric response of metallic perovskites: Sr 1− x La x TiO 3 (51) Baltrusaitis, J.; Mendoza-Sanchez, B.; Fernandez, V.; Veenstra,
(0<∼ x<∼ 0.1). Phys. Rev. B 2001, 63, 113104. R.; Dukstiene, N.; Roberts, A.; Fairley, N. Generalized molybdenum
(29) Fumega, A. O.; Fu, Y.; Pardo, V.; Singh, D. J. Understanding the oxide surface chemical state XPS determination via informed
lattice thermal conductivity of SrTiO 3 from an ab initio perspective. amorphous sample model. Appl. Surf. Sci. 2015, 326, 151−161.
Phys. Rev. Mater. 2020, 4, No. 033606. (52) Wagner, C. D., NIST X-ray photoelectron spectroscopy
(30) Ferrer-Argemi, L.; Yu, Z.; Kim, J.; Myung, N. V.; Lim, J.-H.; Lee, database. NIST Standarad Reference Database 20 2000,
J. Silver content dependent thermal conductivity and thermoelectric DOI: 10.18434/T4T88K.
properties of electrodeposited antimony telluride thin films. Sci. Rep. (53) Choi, J.-G.; Thompson, L. T. XPS study of as-prepared and
2019, 9, 9242. reduced molybdenum oxides. Appl. Surf. Sci. 1996, 93, 143−149.

17031 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

(54) Palacio, C.; Mathieu, H. J. Application of factor analysis to the (78) Liu, J.; Wang, C. L.; Su, W. B.; Wang, H. C.; Li, J. C.; Zhang, J. L.;
AES and XPS study of the oxidation of chromium. Surf. Interface Anal. Mei, L. M. Thermoelectric properties of Sr1− xNdxTiO3 ceramics. J.
1990, 16, 178−182. Alloys Compd. 2010, 117, L54−L56.
(55) Kröger, F.; Vink, H., Relations between the concentrations of (79) Kovalevsky, A. V.; Populoh, S.; Patrício, S. G.; Thiel, P.; Ferro, M.
imperfections in solids. In Solid state physics, Elsevier: 1956; 5, 307− C.; Fagg, D. P.; Frade, J. R; Weidenkaff, A. Design of SrTiO3-based
223, DOI: 10.1016/0022-3697(58)90069-6. thermoelectrics by tungsten substitution. J. Phys. Chem. C 2015, 117,
(56) Anderson, P. W. Absence of diffusion in certain random lattices. 4466−4478.
Phys. Rev. 1958, 109, 1492. (80) Wang, H. C.; Wang, C. L.; Su, W. B.; Liu, J.; Sun, Y.; Peng, H.;
(57) Lee, S.; Bock, J. A.; Trolier-McKinstry, S.; Randall, C. A. Mei, L. M. Doping effect of La and Dy on the thermoelectric properties
Ferroelectric-thermoelectricity and Mott transition of ferroelectric of SrTiO3. J. Am. Ceram. Soc. 2011, 94, 838−842.
oxides with high electronic conductivity. J. Eur. Ceram. Soc. 2012, 32, (81) Li, Y.; Hou, Q.-Y.; Wang, X.-H.; Kang, H.-J.; Yaer, X.; Li, J.-B.;
3971−3988. Wang, T.-M.; Miao, L.; Wang, J. First-principles calculations and high
(58) Ihrig, H.; Hennings, D. Electrical transport properties of n-type thermoelectric performance of La−Nb doped SrTiO 3 ceramics. J.
BaTiO 3. Phys. Rev. B 1978, 17, 4593. Mater. Chem. A 2019, 7, 236−247.
(59) Bernasconi, P.; Biaggio, I.; Zgonik, M.; Günter, P. Anisotropy of (82) Tachibana, M.; Kolodiazhnyi, T.; Takayama-Muromachi, E.
the electron and hole drift mobility in KNbO 3 and BaTiO 3. Phys. Rev. Thermal conductivity of perovskite ferroelectrics. Appl. Phys. Lett.
Lett. 1997, 78, 106. 2008, 93, No. 092902.
(60) Roy, P.; Waghmare, V.; Tanwar, K.; Maiti, T. Large change in
thermopower with temperature driven p−n type conduction switching
in environment friendly Ba x Sr 2− x Ti 0.8 Fe 0.8 Nb 0.4 O 6 double
perovskites. Phys. Chem. Chem. Phys. 2017, 19, 5818−5829.
(61) Saxena, M.; Maiti, T. Metal-like electrical conductivity in La x Sr
2− x TiMoO 6 oxides for high temperature thermoelectric power
generation. Dalton Trans. 2017, 46, 5872−5879.
(62) Maiti, T.; Saxena, M.; Roy, P. Double perovskite (Sr 2 B′ B ″O 6)
oxides for high-temperature thermoelectric power generationA
review. J. Mater. Res. 2019, 34, 107−125.
(63) Austin, I. G.; Mott, N. F. Polarons in crystalline and non-
crystalline materials. Adv. Phys. 1969, 18, 41−102.
(64) Chaikin, P. M.; Beni, G. Thermopower in the correlated hopping
regime. Phys. Rev. B 1976, 13, 647.
(65) Taguchi, H.; Sonoda, M.; Nagao, M. Relationship between angles
for Mn−O−Mn and electrical properties of orthorhombic perovskite-
type (Ca1− xSrx) MnO3. J. Solid State Chem. 1998, 137, 82−86.
(66) Srivastava, D.; Norman, C.; Azough, F.; Schäfer, M. C.;
Guilmeau, E.; Kepaptsoglou, D.; Ramasse, Q. M.; Nicotra, G.; Freer,
R. Tuning the thermoelectric properties of A-site deficient SrTiO 3
ceramics by vacancies and carrier concentration. Phys. Chem. Chem.
Phys. 2016, 18, 26475−26486.
(67) Srivastava, D.; Norman, C.; Azough, F.; Ekren, D.; Chen, K.;
Reece, M.; Kinloch, I. A.; Freer, R. Anisotropy and enhancement of
thermoelectric performance of Sr 0.8 La 0.067 Ti 0.8 Nb 0.2 O 3− δ
ceramics by graphene additions. J. Mater. Chem. A 2019, 7, 24602−
24613.
(68) Cox, P. A., Transition metal oxides: an introduction to their
electronic structure and properties. Oxford university press: 2010,
DOI: 10.1002/ange.19931050352.
(69) Kim, H.-S.; Gibbs, Z. M.; Tang, Y.; Wang, H.; Snyder, G. J.
Characterization of Lorenz number with Seebeck coefficient measure-
ment. APL Mater. 2015, 3, No. 041506.
(70) Cahill, D. G.; Pohl, R. O. Lattice vibrations and heat transport in
crystals and glasses. Annu. Rev. Phys. Chem. 1988, 39, 93−121.
(71) Wang, N.; Li, H.; Ba, Y.; Wang, Y.; Wan, C.; Fujinami, K.;
Koumoto, K. Effects of YSZ additions on thermoelectric properties of
Nb-doped strontium titanate. J. Electron Mater. 2010, 39, 1777−1781.
(72) Callaway, J. Model for lattice thermal conductivity at low
temperatures. Phys. Rev. 1959, 113, 1046.
(73) Slack, G. A. Thermal conductivity of pure and impure silicon,
silicon carbide, and diamond. J. Appl. Phys. 1964, 35, 3460−3466.
(74) Glassbrenner, C. J.; Slack, G. A. Thermal conductivity of silicon
and germanium from 3 K to the melting point. Phys. Rev. 1964, 134,
A1058.
(75) Casimir, H. Note on the conduction of heat in crystals. Physica
1938, 5, 495−500.
(76) Slack, G. A.; Galginaitis, S. Thermal conductivity and phonon
scattering by magnetic impurities in CdTe. Phys. Rev. 1964, 133, A253.
(77) Petersen, A.; Bhattacharya, S.; Tritt, T. M.; Poon, S. J. Critical
analysis of lattice thermal conductivity of half-Heusler alloys using
variations of Callaway model. J. Appl. Phys. 2015, 117, No. 035706.

17032 https://dx.doi.org/10.1021/acssuschemeng.0c03849
ACS Sustainable Chem. Eng. 2020, 8, 17022−17032

You might also like