SnTe is renowned for its promise in advancing energy-related technologies based on thermoelectricity and for its topological crystalline insulator character. Here, we demonstrate that each Mn atom introduces ∼4 μB (Bohr magneton) of magnetic moment to Sn1−xMnxTe. The Curie temperature TC reaches ∼14 K for x = 0.12, as observed in the field dependent hysteresis of magnetization and the anomalous Hall effect. In accordance with a modified two-band electronic Kane model, the light L-valence-band and the heavy Σ-valence-band gradually converge in energy with increasing Mn concentration, leading to a decreasing ordinary Hall coefficient RH and a favorably enhanced Seebeck coefficient S at the same time. With the thermal conductivity κ lowered chiefly via point defects associated with the incorporation of Mn, the strategy of Mn doping also bodes well for efficient thermoelectric applications at elevated temperatures.

SnTe crystallizes in the Fm3¯m(Oh5,No.225) structure with lattice parameter a = 6.318 Å,1,2 well known for thermoelectric (TE) applications.3–16 The band gap, separating the conduction band (CB) and the light valence band (LVB), is located at the L point of the Brillouin zone. Another heavy valence band (HVB), also contributing to transport, is present at the Σ point, with an energy offset ΔE ∼ 0.3–0.4 eV.17,18 The properties of SnTe depend critically on Sn vacancies (VSn, double-hole acceptors), which lead to ∼1020 cm−3 holes in “pristine” SnTe.19 Recently, SnTe has been identified as a topological crystalline insulator (TCI).20,21 Its surface states (SSs) are topologically protected by the crystalline symmetry,22 in contrast to topological insulators (TIs)23,24 where magnetic dopants break the time-reversal symmetry and open up a gap in the SSs.25–27 Introducing magnetism in SnTe is thus of fundamental interest as it may realize phenomena such as the quantum anomalous Hall effect.28–30 Doping SnTe with manganese seems promising, as is known in dilute magnetic semiconductors (DMSs) including II-VI,31,32 III-V,33,34 IV(-VI),35,36 and V2VI3 compounds.37 Although Mn gives rise to ferromagnetism in SnTe, earlier studies have only examined samples with either too low Mn content (≪10 at. %)38 or too high hole concentrations.39 More comprehensive low-temperature magnetic and transport characterizations on a wider range of Mn concentrations are desirable.

Here, we report our recent study on Sn1−xMnxTe (x = 0–0.12), where the Curie temperature TC increases as a function of x. Consistent with a two-band Kane model and a Debye-Callaway thermal conductivity formula, the low-temperature transport properties evolve systematically upon Mn doping, which suggests the convergence of the L and Σ bands, favoring efficient thermoelectricity.

Polycrystalline Sn1−xMnxTe (x = 0–0.12) were synthesized and structurally characterized as detailed in Ref. 8. The X-ray diffraction (XRD) measurements have detected no secondary phases. The samples studied here can be treated as single phase at the XRD level, although transmission electron microscopy (TEM) studies have revealed the presence of a small amount of Mn-rich nanoinclusions for x = 0.12.8 This suggests that the solubility limit of Mn in SnTe is x ∼ 0.12. Rectangular bar-shaped samples (∼1.5 × 2.5 × 5.5 mm3) were cut out at room temperature using a diamond saw. Magnetic and transport measurements were performed over the temperature range of 1.8–300 K. Magnetic studies were conducted using a superconducting quantum interference device (SQUID) magnetometer equipped with a 5.5 T magnet in a Quantum Design Magnetic Property Measurement System (MPMS). The electrical resistivity ρ, Seebeck coefficient S, and thermal conductivity κ were determined using a longitudinal steady-state technique.40 Hall effect and magnetoresistance (MR) measurements were performed using a Linear Research ac bridge with 16 Hz excitation. The uncertainties of ρ, S, and κ are ±3%, ±2%, and ±5%, respectively.

The temperature dependent susceptibilities χm (=M/H per mole) in an applied magnetic field H = 40 Oe (magnetization M depends linearly on H up to ∼50 Oe) are plotted in Fig. 1. With x > 0.03, the system becomes ferromagnetic, manifesting a significant increase of χm below ∼20 K. As shown in the upper inset of Fig. 1, above TC, χm for x ≥ 0.03 follows the Curie-Weiss law

(1)

where C is the Curie constant, T is the absolute temperature, θp is the paramagnetic Curie temperature, and χ0 is a lattice term. The fitting parameters are summarized in Table I, together with the effective moment per Mn atom, peff, deduced from C=NApeff2μB2/3kBx, where NA is the Avogadro's number, μB is the Bohr magneton, and kB is the Boltzmann constant. As exemplified in the lower inset of Fig. 1 for x = 0.12, TC can be determined via the Arrott plot (μ2 vs. −1),41 in reasonable agreement with θp. Future studies are needed to clarify the contribution to the magnetic response from the nanoinclusions at the grain boundaries of x = 0.12. Furthermore, doping beyond the bulk solubility limit x ∼ 0.12, using non-equilibrium technique such as molecular beam epitaxy (MBE), may achieve even higher TC, considering TC of V-doped Sb2Te3 was successfully improved from 22 K (bulk Sb1.97V0.03Te3) to 177 K (Sb1.65V0.35Te3 thin film).42 

FIG. 1.

Zero-field-cooled (ZFC) χm measured with H = 40 Oe for Sn1−xMnxTe. The susceptibilities are well described (upper inset) by the Curie-Weiss law. The θp agrees with TC determined from the Arrott plot, as exemplified for x = 0.12 (lower inset).

FIG. 1.

Zero-field-cooled (ZFC) χm measured with H = 40 Oe for Sn1−xMnxTe. The susceptibilities are well described (upper inset) by the Curie-Weiss law. The θp agrees with TC determined from the Arrott plot, as exemplified for x = 0.12 (lower inset).

Close modal
TABLE I.

Fitting results of Sn1−xMnxTe. The magnetic parameters C (emu K mol−1), θp (K), χ0 (10−5 emu mol−1), peff (per Mn), and TC (K) are from the Curie-Weiss law. The electronic properties ρ (μΩ m), S (μV K−1), RH (10−2 cm3 C−1), μH (cm2 V−1 s−1), Eg (eV),1,52 ΔE (eV), and L (10−8 W Ω K−2) are from experiments or the two-band Kane model at 300 K. The thermal coefficients (μm), A (10−41 s3), and U (10−17 s K−1) are based on the Debye-Callaway formula.

MagneticElectronicThermal
xCθpχ0peffTCρSRHμHEgΔELAU
… … … … … 1.28 6.48 4.49 351 0.18 0.43 2.49 2.12 0.98 0.86 
0.01 … … … … … 1.48 11.9 3.56 241 0.19 0.40 2.59 2.01 1.43 0.87 
0.03 0.066 5.48 −1.62 4.20 1.8 2.43 23.4 2.93 121 0.21 0.36 2.56 0.66 2.41 0.83 
0.07 0.155 7.94 −1.13 4.21 7.5 3.98 42.9 1.33 33.4 0.26 0.14 2.27 0.95 3.83 1.01 
0.12 0.279 15.1 2.74 4.31 14.5 5.19 66.2 0.77 14.8 0.31 −0.10 2.12 1.71 4.75 1.73 
MagneticElectronicThermal
xCθpχ0peffTCρSRHμHEgΔELAU
… … … … … 1.28 6.48 4.49 351 0.18 0.43 2.49 2.12 0.98 0.86 
0.01 … … … … … 1.48 11.9 3.56 241 0.19 0.40 2.59 2.01 1.43 0.87 
0.03 0.066 5.48 −1.62 4.20 1.8 2.43 23.4 2.93 121 0.21 0.36 2.56 0.66 2.41 0.83 
0.07 0.155 7.94 −1.13 4.21 7.5 3.98 42.9 1.33 33.4 0.26 0.14 2.27 0.95 3.83 1.01 
0.12 0.279 15.1 2.74 4.31 14.5 5.19 66.2 0.77 14.8 0.31 −0.10 2.12 1.71 4.75 1.73 

As shown in Fig. 2(a), M of SnTe displays linear field dependence at 1.8 K. With increasing x, hysteresis gradually develops and the high-field saturated M (up to 10 kOe) has quickly reached values exceeding 2000 emu mol−1 for x = 0.12. Measurements at low fields reveal small coercivity fields HC ∼ 10 Oe, indicating the softness of these ferromagnetic semiconductors.

FIG. 2.

(a) The field dependent M for Sn1−xMnxTe at T = 1.8 K showing a hysteresis loop with HC ∼ 10 Oe (inset). (b) The ρyx (T = 5.0 K) of Sn1−xMnxTe develops hysteresis (anomalous Hall effect) for x = 0.07 and 0.12 (lower-right inset). The linearity and the decreasing slope at higher fields indicate a decreasing RH upon Mn doping. The upper-left inset compares the MR in samples with (x = 0.12) and without Mn (x = 0).

FIG. 2.

(a) The field dependent M for Sn1−xMnxTe at T = 1.8 K showing a hysteresis loop with HC ∼ 10 Oe (inset). (b) The ρyx (T = 5.0 K) of Sn1−xMnxTe develops hysteresis (anomalous Hall effect) for x = 0.07 and 0.12 (lower-right inset). The linearity and the decreasing slope at higher fields indicate a decreasing RH upon Mn doping. The upper-left inset compares the MR in samples with (x = 0.12) and without Mn (x = 0).

Close modal

M contributes to the Hall resistivity ρyx via

(2)

where RH is the ordinary Hall coefficient, B (=μ0H) is the magnetic induction, and Rs is the anomalous Hall coefficient. The second term disappears well above TC, restoring linear field dependence for ρyx. The presence of Rs revealed in the hysteretic behavior of ρyx vs. B is one of the most reliable proofs of ferromagnetism. As shown in Fig. 2(b), a distinctly hysteretic character of ρyx is apparent with x > 0.03 at 5.0 K. The very low coercivity [lower inset in Fig. 2(b)] corroborates the M vs. H hysteresis loops in the inset of Fig. 2(a). Considering that M is already well saturated above H ∼ 1 kOe and possesses little field dependence at high fields, the decreasing slope of ρyx in its high-field linear region is mainly due to the decreasing magnitude of RH as x increases. The RH can be extracted from the linear field dependence of ρyx above TC and is displayed in Fig. 3(a). Furthermore, the alignment of spins in Mn-doped samples (e.g., x = 0.12) in the presence of B leads to a negative MR, in contrast to the positive semi-classical parabolic MR of SnTe [upper inset in Fig. 2(b)].

FIG. 3.

The temperature dependent (a) RH, (b) ρ, (c) S, and (d) μH of Sn1−xMnxTe. The inset of (b) shows that ρ of ferromagnetic samples (x = 0.07 and 0.12) develops a peak at ∼TC. The inset of (d) summarizes the band structure (binding energy EB at high symmetry points L and Σ). The LVB (at L) and HVB (at Σ) converge in energy upon Mn doping, leading to a much enhanced S. The numerical fitting (lines) in (a) and (c) simultaneously account for both RH and S within a two-band Kane model. For x = 0.07, the two-band S is compared with contributions from its single-band components SL (from LVB) and SH (from HVB).

FIG. 3.

The temperature dependent (a) RH, (b) ρ, (c) S, and (d) μH of Sn1−xMnxTe. The inset of (b) shows that ρ of ferromagnetic samples (x = 0.07 and 0.12) develops a peak at ∼TC. The inset of (d) summarizes the band structure (binding energy EB at high symmetry points L and Σ). The LVB (at L) and HVB (at Σ) converge in energy upon Mn doping, leading to a much enhanced S. The numerical fitting (lines) in (a) and (c) simultaneously account for both RH and S within a two-band Kane model. For x = 0.07, the two-band S is compared with contributions from its single-band components SL (from LVB) and SH (from HVB).

Close modal

The temperature dependent electronic properties of Sn1−xMnxTe are illustrated in Fig. 3, with 300 K values collected in Table I. Due to the high apparent hole concentration pH (≡1/eRH, e being the elementary charge) ∼1.9× 1020 cm−3 (10 K), ρ of SnTe [Fig. 3(b)] is metallic. While S of SnTe exhibits a nearly linear temperature dependence close to ambient temperatures (dominated by diffusion processes), at low temperatures it develops a broad peak at ∼50 K which originates from strong electron-phonon coupling referred to as the phonon-drag effect.43 Acoustic phonons responsible for this effect also produce a dielectric maximum observed on κ in Fig. 4(a). The phonon drag is substantially washed out with increasing Mn, which strongly scatters acoustic phonons.

FIG. 4.

(a) The temperature dependent κ of Sn1−xMnxTe. (b) The κL is well understood within the Debye-Callaway phonon conductivity model, revealing the enhanced point defect scattering upon Mn doping. The inset in (b) illustrates the dramatically improved ZT at room temperature for the most heavily doped sample, suggesting that efficient TE applications based on these compounds are possible at high temperatures.

FIG. 4.

(a) The temperature dependent κ of Sn1−xMnxTe. (b) The κL is well understood within the Debye-Callaway phonon conductivity model, revealing the enhanced point defect scattering upon Mn doping. The inset in (b) illustrates the dramatically improved ZT at room temperature for the most heavily doped sample, suggesting that efficient TE applications based on these compounds are possible at high temperatures.

Close modal

While the carrier density measured by RH, assuming a single parabolic band, clearly increases with increasing x, the Hall mobility μH (≡RH/ρ) rapidly diminishes [Fig. 3(d)] and is responsible for the rising ρ. Meanwhile, S gradually improves with increasing x [Fig. 3(c)]. This might seem to be a consequence of lowered Fermi level EF in a single band with the density of states D(ε), where ρ and S are proportional to D(ε)−1 and ∂lnD(ε)/∂ε, respectively.44,45 However, such a naïve scenario should be accompanied by a decrease in pH, contradicting its increase as the RH data clearly suggest.

Considering both LVB and HVB [inset of Fig. 3(d)], the overall two-band46–48 electronic properties (RH, S, and the Lorenz number L) are combinations of single-band components

(3)
(4)
(5)

respectively. The mobility ratio b (≡μL/μH) is taken to be 8 for SnTe according to Ref. 18. In a single Kane band with nonparabolicity parameter β (≡kBT/Eg), the number concentration of holes pN is given by

(6)

where md is the density-of-states effective mass and ℏ is the reduced Planck constant. Assuming acoustic phonon dominates,49–51 the generalized Fermi integrals are

(7)

where f0 = (eξη+ 1)−1 is the Fermi-Dirac distribution, ξ = ε/kBT, and η = μc/kBT, with the energy ε and chemical potential μc measured from the band edge and ΔE = kBT (ηL − ηH). The single Kane-band electronic properties are

(8)
(9)
(10)

respectively.

Assuming a Kane-like LVB and a parabolic HVB (βH = 0),3,4,47,48 the two-band Kane model [solid lines in Figs. 3(a), 3(c)] is used to characterize the temperature and composition dependence of RH and S, with a temperature dependent Eg1,52 and constant effective masses (in the unit of rest electron mass me) md,L = 0.1 and md,H = 3, respectively. The fitting parameters are displayed in Table I. A universal shrinking trend (insensitive of Eg and md) of ΔE vs. x (from ∼0.40 eV for SnTe to a somewhat negative value for x = 0.12) has been observed, indicating that Mn gradually changes the band structure and brings into energy convergence the two valence bands–consistent with recent theories.8 An ability to achieve band convergence via chemical doping is beneficial for further improving the TE performance of SnTe,5,53,54 as the temperature-induced convergence in PbTe [ΔE ∼ 0.17 eV (Ref. 55)] is much less feasible in SnTe due to its considerably larger ΔE. As illustrated in Fig. 3(c), contributions to S from the LVB (SL) and HVB (SH) are compared with the total S for x = 0.07. The data suggest that holes are transferred from the LVB to HVB due to the band convergence. Indeed, the less mobile holes in the HVB become more dominant in the electrical transport, leading to an enhanced S and reduced μH.

The κ of Sn1−xMnxTe decreases with increasing x essentially over the whole measured temperature range, as shown in Fig. 4(a). The peak in κ of SnTe is drastically suppressed due to strong phonon scattering by Mn. The lattice thermal conductivity κL, estimated by subtracting the electronic thermal conductivity κe (=LT/ρ, according to the Wiedemann-Franz law) from κ, has been examined using the Debye-Callaway model40 

(11)

where x=ω/kBT, ω is the phonon frequency, v (∼2000 m s−1) is the averaged sound velocity,56,θD (∼177 K) is the Debye temperature,57, is the mean grain size, A represents the Rayleigh point defect scattering, and U denotes the Umklapp processes.58 As shown in Fig. 4(b), the phonon conductivity model describes reasonably well the temperature dependence of κL. However, future effort is needed to improve the model presented here, in order to consider effects from nanoinclusions.59 The fitting parameters are summarized in Table I, revealing a clear trend of enhanced point defect scattering (A) with increasing Mn. The fitted is approximately on par with the values (5–10 μm) typically observed in our SnTe-based polycrystalline samples.8 

The synergistically tuned electrical and thermal transport properties give rise to an enhanced TE figure of merit ZT [≡S2T/ρκ, inset of Fig. 4(b)]. At room temperature, the ZT for x = 0.12 (∼0.1), two orders of magnitude higher than that of SnTe, holds promise for TE performance. Indeed, high ZT values ∼1.3 (900 K) for Sn1−xMnxTe have been recently reported.8–11 

In summary, magnetic and transport studies have been performed on Sn1−xMnxTe, with x approaching the bulk solubility limit ∼0.12. TC of ∼14 K has been achieved for x = 0.12 in the ferromagnetically doped SnTe, as confirmed via the hysteresis in both magnetization and anomalous Hall effect. The study holds promise for further increase of TC using MBE techniques which should enhance Mn incorporation. Ferromagnetic SnTe with high TC may be of interest in future TCI-based (opto-)electronic devices.60 Mn doping also favorably modifies the electronic structure via the L and Σ valence band convergence, and synergistically lowers κ by means of enhanced point defect scattering. This leads to an improved TE performance in comparison to undoped SnTe.

The work was supported by the Revolutionary Materials for Solid State Energy Conversion, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Award No. DE-SC0001054. A part of transport property measurements was performed by H.C. and Q.L. at Brookhaven National Laboratory, which was supported by the U.S. Department of Energy, Office of Basic Energy Science, Materials Sciences and Engineering Division, under Contract No. DE-SC00112704.

1.
J. O.
Dimmock
,
I.
Melngailis
, and
A. J.
Strauss
,
Phys. Rev. Lett.
16
,
1193
(
1966
).
2.
P.
Bauer Pereira
,
I.
Sergueev
,
S.
Gorsse
,
J.
Dadda
,
E.
Müller
, and
R. P.
Hermann
,
Phys. Status Solidi B
250
,
1300
(
2013
).
3.
Q.
Zhang
,
B.
Liao
,
Y.
Lan
,
K.
Lukas
,
W.
Liu
,
K.
Esfarjani
,
C.
Opeil
,
D.
Broido
,
G.
Chen
, and
Z.
Ren
,
Proc. Natl. Acad. Sci. U.S.A.
110
,
13261
(
2013
).
4.
M.
Zhou
,
Z. M.
Gibbs
,
H.
Wang
,
Y.
Han
,
C.
Xin
,
L.
Li
, and
G. J.
Snyder
,
Phys. Chem. Chem. Phys.
16
,
20741
(
2014
).
5.
G.
Tan
,
L.-D.
Zhao
,
F.
Shi
,
J. W.
Doak
,
S.-H.
Lo
,
H.
Sun
,
C.
Wolverton
,
V. P.
Dravid
,
C.
Uher
, and
M. G.
Kanatzidis
,
J. Am. Chem. Soc.
136
,
7006
(
2014
).
6.
G.
Tan
,
F.
Shi
,
J. W.
Doak
,
H.
Sun
,
L.-D.
Zhao
,
P.
Wang
,
C.
Uher
,
C.
Wolverton
,
V. P.
Dravid
, and
M. G.
Kanatzidis
,
Energy Environ. Sci.
8
,
267
(
2015
).
7.
G.
Tan
,
F.
Shi
,
S.
Hao
,
H.
Chi
,
L.-D.
Zhao
,
C.
Uher
,
C.
Wolverton
,
V. P.
Dravid
, and
M. G.
Kanatzidis
,
J. Am. Chem. Soc.
137
,
5100
(
2015
).
8.
G.
Tan
,
F.
Shi
,
S.
Hao
,
H.
Chi
,
T. P.
Bailey
,
L.-D.
Zhao
,
C.
Uher
,
C.
Wolverton
,
V. P.
Dravid
, and
M. G.
Kanatzidis
,
J. Am. Chem. Soc.
137
,
11507
(
2015
).
9.
H.
Wu
,
C.
Chang
,
D.
Feng
,
Y.
Xiao
,
X.
Zhang
,
Y.
Pei
,
L.
Zheng
,
D.
Wu
,
S.
Gong
,
Y.
Chen
 et al.,
Energy Environ. Sci.
8
,
3298
(
2015
).
10.
J.
He
,
X.
Tan
,
J.
Xu
,
G.-Q.
Liu
,
H.
Shao
,
Y.
Fu
,
X.
Wang
,
Z.
Liu
,
J.
Xu
,
H.
Jiang
 et al.,
J. Mater. Chem. A
3
,
19974
(
2015
).
11.
W.
Li
,
Z.
Chen
,
S.
Lin
,
Y.
Chang
,
B.
Ge
,
Y.
Chen
, and
Y.
Pei
,
J. Materiomics
1
,
307
(
2015
).
12.
G.
Tan
,
W. G.
Zeier
,
F.
Shi
,
P.
Wang
,
G. J.
Snyder
,
V. P.
Dravid
, and
M. G.
Kanatzidis
,
Chem. Mater.
27
,
7801
(
2015
).
13.
L.-D.
Zhao
,
X.
Zhang
,
H.
Wu
,
G.
Tan
,
Y.
Pei
,
Y.
Xiao
,
C.
Chang
,
D.
Wu
,
H.
Chi
,
L.
Zheng
 et al.,
J. Am. Chem. Soc.
138
,
2366
(
2016
).
14.
L.-D.
Zhao
,
S.-H.
Lo
,
Y.
Zhang
,
H.
Sun
,
G.
Tan
,
C.
Uher
,
C.
Wolverton
,
V. P.
Dravid
, and
M. G.
Kanatzidis
,
Nature
508
,
373
(
2014
).
15.
L.-D.
Zhao
,
G.
Tan
,
S.
Hao
,
J.
He
,
Y.
Pei
,
H.
Chi
,
H.
Wang
,
S.
Gong
,
H.
Xu
,
V. P.
Dravid
 et al.,
Science
351
,
141
(
2016
).
16.
K.
Peng
,
X.
Lu
,
H.
Zhan
,
S.
Hui
,
X.
Tang
,
G.
Wang
,
J.
Dai
,
C.
Uher
,
G.
Wang
, and
X.
Zhou
,
Energy Environ. Sci.
9
,
454
(
2016
).
17.
R. F.
Brebrick
and
A. J.
Strauss
,
Phys. Rev.
131
,
104
(
1963
).
18.
L. M.
Rogers
,
J. Phys. D: Appl. Phys.
1
,
845
(
1968
).
19.
J. A.
Kafalas
,
R. F.
Brebrick
, and
A. J.
Strauss
,
Appl. Phys. Lett.
4
,
93
(
1964
).
20.
T. H.
Hsieh
,
H.
Lin
,
J.
Liu
,
W.
Duan
,
A.
Bansil
, and
L.
Fu
,
Nat. Commun.
3
,
982
(
2012
).
21.
Y.
Tanaka
,
Z.
Ren
,
T.
Sato
,
K.
Nakayama
,
S.
Souma
,
T.
Takahashi
,
K.
Segawa
, and
Y.
Ando
,
Nat. Phys.
8
,
800
(
2012
).
23.
M. Z.
Hasan
and
C. L.
Kane
,
Rev. Mod. Phys.
82
,
3045
(
2010
).
24.
X.-L.
Qi
and
S.-C.
Zhang
,
Rev. Mod. Phys.
83
,
1057
(
2011
).
25.
Y. L.
Chen
,
J.-H.
Chu
,
J. G.
Analytis
,
Z. K.
Liu
,
K.
Igarashi
,
H.-H.
Kuo
,
X. L.
Qi
,
S. K.
Mo
,
R. G.
Moore
,
D. H.
Lu
 et al.,
Science
329
,
659
(
2010
).
26.
J. G.
Checkelsky
,
J.
Ye
,
Y.
Onose
,
Y.
Iwasa
, and
Y.
Tokura
,
Nat. Phys.
8
,
729
(
2012
).
27.
I.
Lee
,
C. K.
Kim
,
J.
Lee
,
S. J. L.
Billinge
,
R.
Zhong
,
J. A.
Schneeloch
,
T.
Liu
,
T.
Valla
,
J. M.
Tranquada
,
G.
Gu
 et al.,
Proc. Natl. Acad. Sci. U.S.A.
112
,
1316
(
2015
).
28.
C.-Z.
Chang
,
J.
Zhang
,
X.
Feng
,
J.
Shen
,
Z.
Zhang
,
M.
Guo
,
K.
Li
,
Y.
Ou
,
P.
Wei
,
L.-L.
Wang
 et al.,
Science
340
,
167
(
2013
).
29.
J. G.
Checkelsky
,
R.
Yoshimi
,
A.
Tsukazaki
,
K. S.
Takahashi
,
Y.
Kozuka
,
J.
Falson
,
M.
Kawasaki
, and
Y.
Tokura
,
Nat. Phys.
10
,
731
(
2014
).
30.
C.
Fang
,
M. J.
Gilbert
, and
B. A.
Bernevig
,
Phys. Rev. Lett.
112
,
046801
(
2014
).
31.
A.
Haury
,
A.
Wasiela
,
A.
Arnoult
,
J.
Cibert
,
S.
Tatarenko
,
T.
Dietl
, and
Y.
Merle d'Aubigné
,
Phys. Rev. Lett.
79
,
511
(
1997
).
32.
P.
Sharma
,
A.
Gupta
,
K. V.
Rao
,
F. J.
Owens
,
R.
Sharma
,
R.
Ahuja
,
J. M. O.
Guillen
,
B.
Johansson
, and
G. A.
Gehring
,
Nat. Mater.
2
,
673
(
2003
).
33.
H.
Ohno
,
A.
Shen
,
F.
Matsukura
,
A.
Oiwa
,
A.
Endo
,
S.
Katsumoto
, and
Y.
Iye
,
Appl. Phys. Lett.
69
,
363
(
1996
).
34.
M. E.
Overberg
,
C. R.
Abernathy
,
S. J.
Pearton
,
N. A.
Theodoropoulou
,
K. T.
McCarthy
, and
A. F.
Hebard
,
Appl. Phys. Lett.
79
,
1312
(
2001
).
35.
T.
Story
,
R. R.
Gała̧zka
,
R. B.
Frankel
, and
P. A.
Wolff
,
Phys. Rev. Lett.
56
,
777
(
1986
).
36.
Y. D.
Park
,
A. T.
Hanbicki
,
S. C.
Erwin
,
C. S.
Hellberg
,
J. M.
Sullivan
,
J. E.
Mattson
,
T. F.
Ambrose
,
A.
Wilson
,
G.
Spanos
, and
B. T.
Jonker
,
Science
295
,
651
(
2002
).
37.
Y. S.
Hor
,
P.
Roushan
,
H.
Beidenkopf
,
J.
Seo
,
D.
Qu
,
J. G.
Checkelsky
,
L. A.
Wray
,
D.
Hsieh
,
Y.
Xia
,
S. Y.
Xu
 et al.,
Phys. Rev. B
81
,
195203
(
2010
).
38.
M.
Inoue
,
K.
Ishii
, and
H.
Yagi
,
J. Phys. Soc. Jpn.
43
,
903
(
1977
).
39.
C. W. H. M.
Vennix
,
E.
Frikkee
,
P. J. T.
Eggenkamp
,
H. J. M.
Swagten
,
K.
Kopinga
, and
W. J. M.
de Jonge
,
Phys. Rev. B
48
,
3770
(
1993
).
40.
H.
Chi
,
W.
Liu
,
K.
Sun
,
X.
Su
,
G.
Wang
,
P.
Lošt'ák
,
V.
Kucek
,
Č.
Drašar
, and
C.
Uher
,
Phys. Rev. B
88
,
045202
(
2013
).
41.
42.
Z.
Zhou
,
Y.-J.
Chien
, and
C.
Uher
,
Appl. Phys. Lett.
87
,
112503
(
2005
).
43.
G.
Wang
,
L.
Endicott
,
H.
Chi
,
P.
Lošt'ák
, and
C.
Uher
,
Phys. Rev. Lett.
111
,
046803
(
2013
).
44.
G. J.
Snyder
and
E. S.
Toberer
,
Nat. Mater.
7
,
105
(
2008
).
45.
J. P.
Heremans
,
V.
Jovovic
,
E. S.
Toberer
,
A.
Saramat
,
K.
Kurosaki
,
A.
Charoenphakdee
,
S.
Yamanaka
, and
G. J.
Snyder
,
Science
321
,
554
(
2008
).
46.
H. J.
Goldsmid
,
Electronic Refrigeration
(
Pion
,
London
,
1986
).
47.
H.
Sun
,
X.
Lu
,
H.
Chi
,
D. T.
Morelli
, and
C.
Uher
,
Phys. Chem. Chem. Phys.
16
,
15570
(
2014
).
48.
T. C.
Chasapis
,
Y.
Lee
,
E.
Hatzikraniotis
,
K. M.
Paraskevopoulos
,
H.
Chi
,
C.
Uher
, and
M. G.
Kanatzidis
,
Phys. Rev. B
91
,
085207
(
2015
).
49.
Y. I.
Ravich
,
B. A.
Efimova
, and
I. A.
Smirnov
,
Semiconducting Lead Chalcogenides
(
Plenum Press
,
New York
,
1970
).
50.
D. I.
Bilc
,
S. D.
Mahanti
, and
M. G.
Kanatzidis
,
Phys. Rev. B
74
,
125202
(
2006
).
51.
H.
Wang
,
E.
Schechtel
,
Y. Z.
Pei
, and
G. J.
Snyder
,
Adv. Energy Mater.
3
,
488
(
2013
).
52.
S.-H.
Wei
and
A.
Zunger
,
Phys. Rev. B
35
,
2340
(
1987
).
53.
Y. Z.
Pei
,
X. Y.
Shi
,
A.
LaLonde
,
H.
Wang
,
L. D.
Chen
, and
G. J.
Snyder
,
Nature
473
,
66
(
2011
).
54.
W.
Liu
,
X.
Tan
,
K.
Yin
,
H.
Liu
,
X.
Tang
,
J.
Shi
,
Q.
Zhang
, and
C.
Uher
,
Phys. Rev. Lett.
108
,
166601
(
2012
).
55.
H.
Sitter
,
K.
Lischka
, and
H.
Heinrich
,
Phys. Rev. B
16
,
680
(
1977
).
56.
T.
Seddon
,
S. C.
Gupta
, and
G. A.
Saunders
,
Solid State Commun.
20
,
69
(
1976
).
57.
E. R.
Cowley
,
J. K.
Darby
, and
G. S.
Pawley
,
J. Phys. C: Solid State Phys.
2
,
1916
(
1969
).
58.
C. J.
Glassbrenner
and
G. A.
Slack
,
Phys. Rev.
134
,
A1058
(
1964
).
59.
H.
Zhang
and
A. J.
Minnich
,
Sci. Rep.
5
,
8995
(
2015
).
60.
J.
Shen
and
J. J.
Cha
,
Nanoscale
6
,
14133
(
2014
).