Sei sulla pagina 1di 64

Progress in Materials Science 97 (2018) 283–346

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

High-performance SnSe thermoelectric materials: Progress and


T
future challenge
Zhi-Gang Chena,b,⁎,1, Xiaolei Shib,1, Li-Dong Zhaoc, Jin Zoub,d,⁎
a
Centre for Future Materials, University of Southern Queensland, Springfield, QLD 4300, Australia
b
Materials Engineering, The University of Queensland, St Lucia, QLD 4072, Australia
c
School of Materials Science and Engineering, Beihang University, Beijing 100191, China
d
Centre for Microscopy and Microanalysis, The University of Queensland, St Lucia, QLD 4072, Australia

A R T IC LE I N F O ABS TRA CT

Keywords: Thermoelectric materials offer an alternative opportunity to tackle the energy crisis and en-
Thermoelectric materials vironmental problems by enabling the direct solid-state energy conversion. As a promising
SnSe candidate with full potentials for the next generation thermoelectrics, tin selenide (SnSe) and its
Electrical conductivity associated thermoelectric materials have been attracted extensive attentions due to their ultralow
Seebeck coefficient
thermal conductivity and high electrical transport performance (power factor). To provide a
Thermal conductivity
thorough overview of recent advances in SnSe-based thermoelectric materials that have been
revealed as promising thermoelectric materials since 2014, here, we first focus on the inherent
relationship between the structural characteristics and the supreme thermoelectric performance
of SnSe, including the thermodynamics, crystal structures, and electronic structures. The effects
of phonon scattering, pressure or strain, and oxidation behavior on the thermoelectric perfor-
mance of SnSe are discussed in detail. Besides, we summarize the current theoretical calculations
to predict and understand the thermoelectric performance of SnSe, and provide a comprehensive
summary on the current synthesis, characterization, and thermoelectric performance of both
SnSe crystals and polycrystals, and their associated materials. In the end, we point out the
controversies, challenges and strategies toward future enhancements of the SnSe thermoelectric
materials.

1. Introduction

Thermoelectric materials, enabling the direct solid-state energy conversion between heat and electricity [1,2], offer an alternative
solution with full potentials for power generation and refrigeration [3–6]. In order to evaluate the thermoelectric efficiency, the
figure of merit (ZT) is used and is defined as:
S 2σ S 2σ
ZT = T= T,
κ κ e + κl (1)
where S, σ, κ, κl, κe, and T are the Seebeck coefficient, electrical conductivity, thermal conductivity, lattice thermal conductivity,
electronic thermal conductivity, and absolute temperature, respectively [7–9]. The power factor (S2σ) is associated with electrical


Corresponding authors at: Centre for Future Materials, University of Southern Queensland, Springfield, QLD 4300, Australia (ZGC); Materials Engineering, The
University of Queensland, St Lucia, QLD 4072, Australia (JZ).
E-mail addresses: zhigang.chen@usq.edu.au (Z.-G. Chen), j.zou@uq.edu.au (J. Zou).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.pmatsci.2018.04.005
Received 20 September 2017; Received in revised form 21 March 2018; Accepted 24 April 2018
Available online 30 April 2018
0079-6425/ © 2018 Elsevier Ltd. All rights reserved.
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a Single crystal SnSe doped with 6% Bi


Half-Heuslers Ti0.5Zr0.25Hf0.25NiSn
2.0
Mg2Si0.5Sn0.5
Single crystal In4Se2.35
1.5 Skutterudites Ba0.08La0.05Yb0.04Co4Sb12
Clathrates Ba9Ga16Ge30
ZT
1.0 Si80Ge20
Bi2Te2.7Se0.3
0.5 La3Te3.35Sb0.65
(PbTe)0.84(PbSe)0.07(PbS)0.07
0.0 Zintls Mg3Sb1.48Bi0.48Te0.04
Mg2Sn0.75Ge0.25
400 600 800 1000 1200

T (K)

b 2.5

2.0

1.5
ZT

1.0

0.5

0.0

400 600 800 1000 1200

T (K)
Fig. 1. Temperature-dependent ZT values for some typical classes of (a) n-type [42–53] and (b) p-type [6,15,54–63] bulk thermoelectric materials.

transports [10]. As shown in Eq. (1), a relative high ZT requires a high S2σ and/or a low κ [11,12]. So far, significant progress has
been made in enhancing ZT through increasing S2σ by resonant state doping [13–16], minority carrier blocking [17–20], band
convergence [21–24], quantum confinement [25–28]; and/or reducing κ by nanostructuring [29–32], hierarchical architecturing
[6,33–35], and matrix with nanoprecipitates [36–39]. In general, one thermoelectric component contains an n-type and a p-type
thermoelectric material, so that the developments of both n-type and p-type high-performance thermoelectric materials are necessary
for the practical use of thermoelectric devices [40,41]. Fig. 1(a) and (b) present the major milestones achieved for ZT values as a
function of T both for n-type [42–53] and p-type [6,15,54–63] bulk thermoelectric materials, in which a remarkable high ZT of ∼2.6
at 923 K found in the p-type single crystal SnSe (along the b-axis) was obtained as so-far the world record [61], as well as a relative
high ZT of 2.2 at 773 K obtained from the n-type Bi-doped single crystal SnSe (also along the b-axis) [52]. Such outstanding ther-
moelectric performances were derived from the ultralow κ observed from the high temperature SnSe phase at T > 973 K. Inspired by
the ultralow κ and high electrical performance (σ, S, and in turn S2σ), SnSe-based thermoelectric materials have attracted significant
attention in recent years [64], and shortly became a globally focused research topic. Although SnSe single crystals have demonstrated
outstanding thermoelectric property, they still have their own limitations for practical applications, mainly due to their poor me-
chanical properties, rigid-controlled crystal growth conditions, and high production cost for industrial scale-up [65]. For this reason,
the development of high performance polycrystalline SnSe became a research focus, since there is a significant potential for en-
hancing the ZT values of polycrystalline SnSe materials through systematic optimization, such as texturing [66], doping [67], and
alloying [64]. So far, the ZT values of polycrystalline SnSe materials have been improved from 0.5 to nearly 1.7 [68]. However, it
should be noted that the thermoelectric performance of polycrystalline SnSe is still much lower than its single crystal counterpart due
to its comparatively high κ and low σ [69]. Achieving an ultralow κ to be comparable to the single crystal counterpart has been a
challenge and needs strategies. In this review, we aim to provide a thorough summary of current research on structural character-
istics, theoretical calculations, syntheses, characterizations, and thermoelectric performance of both single-crystal and polycrystalline
SnSe and its associated materials. On this basis, we also discuss the controversies, challenges and strategies toward future en-
hancements of the thermoelectric performance of SnSe-based materials.

284
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

2. Structural characteristics of SnSe

2.1. Thermodynamics

SnSe is an inorganic compound semiconductor [70], that has a Molar mass of 197.67 g mol−1, a theoretical density of
6.179 g cm−3 at room temperature [71], and a melting point of 1134 K [72,73]. To obtain a stable phase of SnSe compound during
the synthesis process, it is essential to understand the thermodynamic data of SnSe.

2.1.1. Formulae and relevant enthalpies


Sn can directly react with Se to form stable SnSe compounds. In a mass-spectrometric investigation of the vapor in equilibrium
with solid SnSe, a number of reaction enthalpies were measured [74]. Taking the thermochemical data into account, the fundamental
reaction formulae and relevant enthalpies between Sn, Se and SnSe can be summarized as follows [74,75]:
o −1
Se(s ) + Sn(s ) → SnSe(s ) ΔH298,f (SnSe) = −21.5 ± 1.7 kcal mol ;

SnSe(s ) → Sn(s ) + Se(s ) ΔH = 16.5 ± 2.0 kcal mol−1;

SnSe(g ) → Sn(g ) + Se(g ) ΔH = 85.5 ± 2.0 kcal mol−1;

SnSe(g ) → SnSe(s ) ΔH = −52.4 ± 1.0 kcal mol−1;

Se(s ) → Se(g ) ΔH = 49.4 kcal mol−1; and

Sn(s ) → Sn(g ) ΔH = 72.0 ± 2.0 kcal mol−1.

These formulae are essential for exploring the nature of the atomic and molecular species composing the saturated vapor of SnSe.
The total pressure of SnSe has been measured using classical techniques [75]. The evaporation was carried out below the melting
point (1134 K). For the dissociation energy of SnSe, Do″ (SnSe) = 4.54 eV [74]. Besides, the Gibbs formation energy of SnSe was
determined by an electromotive force (emf) measurement, in the temperature range from 435 to 485 K [76], and the value of Δf G 0 =
−95389 − 3.184 T J mol−1. The heat capacity (Cp) of SnSe in both temperature ranges from 373 to 1135 K and from 233 to 573 K
was measured [61,77], which is important to determine the stability of the SnSe compounds. Besides, the thermodynamic data can be
used as the guidance for designing their fabrications, especially for the growth of SnSe crystals via the Bridgman method [61,78–80]
or solvothermal method [81–86].

2.1.2. Sn-Se binary phase diagram


The binary phase diagram is very useful to provide the thermodynamic information. Fig. 2(a) is the Sn-Se binary phase diagram
[72,87], in which two intermediate phases can be found as SnSe and SnSe2. In the Sn-SnSe region, a liquid miscibility gap, a
monotectic reaction, and a eutectic reaction can be observed. Besides, eutectic reactions can also be found in the SnSe-SnSe2 and
SnSe2-Se regions. The two eutectic points, L1 ⇔ (Sn) + SnSe and L2 ⇔ SnSe2 + (Se), are very close to pure Sn and pure Se,
respectively. For the intermediate phase of SnSe, there are two stable SnSe phases, denoted as α-SnSe (low temperature) and β-SnSe
(high temperature), respectively.
As an important and convenient synthesis method for fabricating SnSe, melting routes are widely used [65]. Generally, in order to
obtain high pure and stable SnSe compounds, the melting route is to set T > 1134 K – the melting point [65,66,88–90]. To pursue a
high density and stable structure in the sintering process of polycrystalline SnSe, it is generally sintered at a high temperature ranging
from 800 to 1105 K [91]. Therefore, the sintered products are often β-SnSe phase.

2.1.3. Thermogravimetric analysis


To investigate the thermal ability of SnSe, Fig. 2(b) and (c) plots the temperature-dependent thermogravimetric analysis (TGA)
and differential thermal analysis (DTA) curves of SnSe in the temperature range from 300 to 1273 K, performed in an inert nitrogen
atmosphere [92]. The TGA curve shows a slight increase of weight when temperature reaches ∼853 K and a peak at ∼993 K due to
the nitrogen adsorption on the surface during the analysis [92]. Such an adsorption is supported by an endothermic peak at ∼993 K
found in the DTA curve. Then a continuous loss of weight in TGA from ∼993 to ∼1133 K indicates a volatilization of Se [92]. This
suggests that the volatilization of SnSe starts at the temperature much lower than its melting point. The DTA curve shows a sharp
exothermic peak accompanied by an endothermic peak, which can be attributed to a thermally induced phase transition of SnSe
before melting at ∼1133 K [92]. This information suggests that α-SnSe is much more stable than β-SnSe and β-SnSe is easily to crack
at high temperature [68]. There is no weight change seen in the TGA curve from 1133 K to 1273 K. However, in the DTA curve, there
is an endothermic peak in this temperature range (at 1253 K), suggesting the formation of Sn–N–Se phase which may be formed due
to the incorporation of nitrogen into the sample. However, it is necessary to carry out X-ray diffraction studies at this temperature to
confirm the presence of this Sn–N–Se phase.
The TGA and DTA of SnSe are important indicatives for the potential use of SnSe as a thermoelectric material in the device for the
real application. Considering the instability of SnSe at high temperature, SnSe-based thermoelectric materials should be practically
used for T < 1000 K.

285
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

b 105 c 0.03
T Difference (K mg-1)

100 0.00
95
Weight (%)

-0.03
90
-SnSe -SnSe
-0.06
85 -SnSe -SnSe
80 -0.09
TGA curve DTA curve
75 -0.12
300 450 600 750 900 1050 1200 300 450 600 750 900 1050 1200
T (K) T (K)
Fig. 2. (a) Sn-Se binary phase diagram [72,87], reproduced and revised with permission, Copyright 2006, ASM International, (b) TGA curve and (c)
DTA curve for SnSe taken in nitrogen atmosphere. Reproduced with permission [92]. Copyright 2007, Elsevier Publishing.

2.2. Crystal structure

As mentioned above, SnSe has two stable phases: α-SnSe (T < ∼800 K) and β-SnSe (T > ∼800 K). These two phases possess a
similar crystal structure (orthogonal), but different lattice parameters and space groups, as summarized in Table 1, where S.D. is the
abbreviation of Strukturbericht designation [72,87].

2.2.1. α-SnSe and β-SnSe


Low-temperature α-SnSe (T < ∼800 K) has an orthogonal structure with lattice parameters of a = 11.37 Å, b = 4.19 Å, and
c = 4.44 Å and a space group of Pnma (#62). Fig. 3(a) shows the atomic structure of α-SnSe in a unit cell, and Fig. 3(b)–(d) show the
projected crystal structure of α-SnSe viewed along the a-, b-, and c-axes, respectively. As can been seen, α-SnSe possesses a typical
double-layered structure [93], which shares the same structure with black phosphorus [94]. α-SnSe consists of eight atoms in a unit

Table 1
Crystal structure data of SnSe system [72,87].
Phase Composition, % Se Pearson symbol Space group S. D. Lattice Parameters (Å) Prototype

α-SnSe 50 oP8 Pnma B16 a = 11.37, b = 4.19, c = 4.44 GeS


β-SnSe 50 oC8 Cmcm Bf a = 4.31, b = 11.71, c = 4.32 CrB

286
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 3. (a) Unit cell of α-SnSe and crystal structure along (b) a-axis, (c) b-axis, and (d) c-axis.

cell [95]. The Sn and Se atoms are joined with strong heteropolar bonds to form the crystalline layers, consisting of two planes of
zigzag Sn-Se type chains [95,96]. Each Sn atom is bonded to three neighboring Se atoms and each Se atom is also bonded to three
neighboring Sn atom. The adjacent layers are mainly bound to each other with a combination of van der Waals forces and a long-
range electrostatic attractions [95,97].
High temperature β-SnSe (∼800 K < T < 1134 K) has an orthogonal structure with lattice parameters of a = 4.31 Å, b = 11.71 Å,
and c = 4.42 Å and a space group of Cmcm (#63) for [98]. Fig. 4(a) shows the atomic structure of β-SnSe in a unit cell, and
Fig. 4(b)–(d) show the projected crystal structure of β-SnSe viewed along the a-, b-, and c-axes, respectively. Similar to α-SnSe, β-SnSe
has a typical double-layered structure. Sn and Se atoms are joined with strong heteropolar bonds. Different from α-SnSe, β-SnSe has a
higher crystal symmetry. When cooling from β-SnSe to α-SnSe, a phase transition occurs at ∼800 K, as well as a transformation of
crystallographic axes.

2.2.2. Variation of parameters


To illustrate the change of atom position of Sn and Se during the phase transition, Fig. 5(a) and (b) show the variation of the
atomic positional parameters (x and y) of Sn and Se atoms [71]. The y parameters are practically constant with temperature, whereas
the x parameters change continuously with temperature and become equal to those of the β-SnSe phase. Therefore, the parameter x is
identified as the primary order parameter [71]. Fig. 5(c) and (d) shows the variation of the thermal parameters of the Sn and Se atoms
with increasing the temperature [71]. At room temperature the components U11, U22 and U33 of the Sn and Se atoms are practically
the same. When the temperature increase, the anisotropy in the thermal parameters becomes obvious. In the case of the Sn atoms, the
component U11 becomes much larger than U22 and U33. In the case of Se atoms, U11 and U22 becomes larger than U33.
Because of the change of crystallographic parameters and thermal parameters via phase transition, the α-SnSe phase expands its
volume by ∼2.5% during cooling from β-SnSe. Such a volume expansion is the main reason causing the breakage of silica tube and
sample damage during crystal growth of SnSe [68].

2.3. Anharmonic bonding

As a property of lattice vibrations, anharmonicity affects the heat transportation in SnSe structure. The anharmonic bonding
between Sn and Se is one of the most important features of the SnSe crystal structure, which is the key factor for the obtained ultralow
κ of SnSe [99].

287
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 4. (a) Unit cell of β-SnSe and crystal structure along (b) a-axis, (c) b-axis, and (d) c-axis.

2.3.1. Harmonicity and anharmonicity


To illustrate the anharmonicity in the SnSe crystal structure, Fig. 6(a) shows the relationship between harmonicity and anhar-
monicity [68,100], where Ф(r), a0 and r are the potential energy, lattice parameter, and distance between two adjacent atoms,
respectively. For harmonicity, during the transport of a phonon, if an atom is forced to deviate from its equilibrium position, the
applied force is proportional to its displacement [100]. The proportionality of this relationship is referred as the spring constant (or
stiffness) [100]. The harmonicity shows a balanced phonon transport [100]. On the other hand, anharmonicity represents that the
applied force is no longer proportional to the displacement when an atom is under a force. As shown in Fig. 6(a), the anharmonicity is
the deviation from a linear relationship between the displacement and restoring force, showing an imbalanced phonon transport
[101,102]. High anharmonicity therefore results in enhanced phonon – phonon scattering and in turn reduces κl.

2.3.2. Grüneisen parameter


It should be noted that in real materials, all bonds are anharmonic, but the degree of anharmonicity varies from different ma-
terials [22,100,103–106]. To verify the strength of anharmonicity, the Grüneisen parameter (γ) is used, which can be described as
follows [107]:

3βBVm
γ= ,
Cv (2)

where β is the volume thermal expansion coefficient, B is the isothermal bulk modulus, Vm is the molar volume, and Cv is the isochoric
specific heat per mol, respectively. γ = 0 when the crystal is fully harmonic bonded, and large γ value represents a strong phonon
scattering. For SnSe, the average γ values of SnSe along a-, b- and c-axes are 4.1, 2.1, 2.3, respectively [61]. As can be seen, the γ
values along the three different axes are different, indicating that SnSe possesses strong anharmonicity, which is the key reason for
giving rise to an ultralow κ of SnSe, especially along the a-axis [64,102]. Besides, the large γ values of SnSe reflect the unique crystal
structure. It was reported that by using the Grüneisen theory, first-principles calculations indicated that bulk SnSe possesses negative
Possion’s ratio, similar to other two-dimensional (2D) systems, such as black phosphorus [108], 2D SnSe and phosphorene [109].
Fig. 6(b) illustrates the resonant bonding of the highly distorted SnSe7 polyhedra caused by the long pair of Sn2+ [68]. In the
SnSe7 polyhedra, one Sn atom is surrounded by seven Se atoms, which exists unbalanced forces around the Sn atom caused by four
long Sn-Se bonds and three short Sn-Se bonds. In this regard, SnSe possesses a soft lattice. Along b- and c-axes, the Sn-Se bond length
cannot change directly if the mechanical stress was applied. Along the a-axis, the weaker bonding between SnSe layers can provide a
good stress buffer, thus dissipating phonon transport laterally [100,110]. Such a unique crystal structure may induce an ultralow κl

288
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 0.15 b 0.87 Se
Sn
0.12
0.84 -SnSe
0.09 -SnSe
x, y

x, y
x -SnSe
0.06 0.51 y
-SnSe
0.03 x 0.48
y
0.00
0.45
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
c 0.09 -SnSe d 0.06 -SnSe
U11 U11
U22 Sn Se
U22
0.06 U33 0.04 U33

Uij (A2)
Uij (A2)

o o

0.03
0.02
-SnSe -SnSe
0.00
0.00
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
Fig. 5. Temperature-dependent positional parameters x (green line) and y (red line) of (a) Sn and (b) Se atoms, and temperature-dependent thermal
parameters of (c) Sn and (d) Se atoms. Reproduced with permission [71]. Copyright 1986, Elsevier Publishing.

Fig. 6. (a) Schematic diagrams of harmonicity and anharmonicity and (b) resonant bonding in SnSe. Here Ф(r), a0 and r are the potential energy,
lattice parameter, and distance between two adjacent atoms, respectively. Reproduced with permission [68]. Copyright 2016, RSC Publishing.

along the a-axis. Other studies showed that via heat capacity measurements, two characteristic vibrational energy scales in SnSe were
revealed, which were corresponding to hard and soft substructures [111]. These distinct substructures derived from the strong bond
divergence in SnSe, and the soft substructure was largely responsible for the thermal transport, which was consistent with the strong
Umklapp scattering and low κ values observed in SnSe [111].

2.3.3. Lattice dynamics


To understand the lattice dynamics of SnSe, Fig. 7(a) and (b) plot the phonon dispersion and density of states (DOS) of the
optimized equilibrium α-SnSe and β-SnSe structures through a first-principles lattice-dynamics study, respectively [112]. As can be

289
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 7. Phonon dispersions and density of states (DOS) of the optimized equilibrium (a) α-SnSe and (b) β-SnSe, (c) comparison of the DOS curves,
and (d) cumulative percentage of states under each curve as a function of frequency [112].

seen, compared to α-SnSe, β-SnSe displays a phonon softening and enhanced three-phonon scattering, leading to an anharmonic
damping of the low-frequency modes and hence the thermal transport [112,113]. Fig. 7(c) compares the DOS curves for the equi-
librium α-SnSe and β-SnSe structures. From which, an obvious red shift of the higher-frequency modes (with frequencies over 3 THz)
is observed in β-SnSe with respect to α-SnSe [112]. To quantify this shift, Fig. 7(d) plots the cumulative percentage of states under
each curve as a function of frequency [112]. There are a similar number of states below 3 THz in both phases, but between 3 and 4
THz there are ∼15% more states in β-SnSe than in α-SnSe [112].
A combination of inelastic neutron scattering measurements and first-principles simulations was performed to illustrate that the
strong anharmonicity of lattice dynamics results from a Jahn-Teller electronic instability, which is driven by the lattice distortion in
SnSe [114]. The phonons in SnSe are strongly anisotropic, associated with a soft-mode lattice instability driven by the resonantly
bonding Se p-states and stereochemically active Sn lone pair [115]. Comprehensive first-principles calculations indicated that the
remarkable anisotropy (κlb > κlc > κla ) in the κl is found in SnSe, which is mainly due to the anisotropic group velocity caused by the
layered lattice structure [110]. The phonon relaxation times also indicate high anharmonicity in SnSe, accounting for the extremely
low κl [116].

2.4. Phonon scattering effect

As discussed above, the ultralow κ of SnSe mainly derives from the layered crystal structure and strongly anharmonic bonding
[117]. However, phonon scatterings also frequently occur at grain boundaries and lattice defects, such as point defects and dis-
locations [65,81,118,119], especially in polycrystalline materials. Phonon scattering effects have been used as powerful tools to
strengthen the scattering of phonons during transportation, in turn further reduce κl [31,48,120–124]. Fig. 8 illustrates the main
potential phonon scattering effects occurring in polycrystalline SnSe materials. In polycrystalline SnSe materials, potential phonon
scattering sources include point defects, dislocations, precipitates, and grain boundaries [125].

2.4.1. Point defects


Point defects (include atom vacancies and substitutions) belong to atomic scale, dislocations and precipitates belong to nano
scale, and grain boundaries belong to sub-micro scale [125]. Fig. 9(a) shows a Sn vacancy in SnSe structure characterized by STM
[52], which illustrates that there are Sn vacancies exist in the SnSe structure, making the whole system presenting p-type [126].
To understand the phonon scattering effect by point defects, combined experimental measurements and computational modelling
showed that the induced extended strain fields on point defects result in extra phonon scattering in SnSe-based thermoelectric
materials [127]. The strength and extent of the induced strain fields depend strongly upon defect chemistry [127]. Besides, the
existence of strain fields can obviously distort the SnSe crystal structure and further enhance the anharmonicity [127]. By alloying

290
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 8. Schematic diagrams of phonon scattering effects occur in SnSe.

d 0.25 Umklapp (U)

Grain Boundary (B) +


0.20
(f) (pW sm-1K-1)

Point Defect (P)

0.15
Dislocation (D)
0.10
U
s

0.05 U+B+P
U+B+P+D
0.00
0.0 0.3 0.6 0.9 1.2 1.5 1.8
Frequency (THz)

Fig. 9. (a) One point defect in SnSe characterized by STM, reproduced with permission [52], Copyright 2016, Nature Publishing Group, (b) HRTEM
image showing dislocation cores, (c) EBSD result of polycrystalline SnSe, reproduced with permission [84], Copyright 2016, ACS Publications, (d)
Illustration of full-spectrum phonon scattering, reproduced with permission [132], Copyright 2015, American Association for the Advancement of
Science, (e) low magnification and (f) HRTEM images of SnSe matrix with precipitates along [1 0 0], inset figures in panels (e) and (f) are the FFT
images in the matrix and precipitates, respectively. Reproduced with permission [67], Copyright 2016, John Wiley and Sons.

291
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

with barium (Ba), the strain fields cause large shifts in the equilibrium positions of atoms, which is up to 0.5 Å [127]. Thus, in-
troducing point defects was an effective method for the design and optimization of SnSe-based thermoelectric materials
[110,128,129].

2.4.2. Dislocations and grain boundaries


Apart from the point defect scattering, dislocation and grain boundary scatterings also have obvious influence on reducing κ
[125,130,131]. Fig. 9(b) shows a high resolution transmission electron microscopy (HRTEM) image of dislocations in SnSe, and
Fig. 9(c) shows the electron back-scattered diffraction (EBSD) result of polycrystalline SnSe to show grain boundaries and grains with
different orientations [84]. These dislocations and grain boundaries can induce extended strain fields and in turn result in strong
phonon scatterings [125].
To compare the ability of scattering phonons between point defects, dislocations and grain boundaries, Fig. 9(d) illustrates the
full-spectrum phonon scattering by different sources [132]. The phonons cover a broad spectrum of frequency (f), and κl can be
expressed as a sum of contributions from different frequencies [133,134]:

κl = ∫ κs (f ) df , (3)

where κs(f) is the spectral lattice thermal conductivity [132], and can be described as:
κs (f ) = Cp (f ) × v 2 (f ) × τ (f ), (4)

where Cp(f) is the spectral heat capacity of phonons, v(f) is the phonon velocity, and τ(f) is the scattering time. It is well known that
phonons in all crystalline materials are scattered by other phonons through Umklapp scattering [132], follow the relationship of
τU−1∼f2. The Debye approximation for phonons has a relationship of Cp(f)∼f2. Combining this with the former one gives
κs(f) = constant. This leaves a wide range of phonon frequencies where all frequencies contribute to κ. Point defects scatter mostly
high-frequency phonons (τP−1∼f4), and grain boundaries scatter mostly low-frequency phonons (τB−1∼f0) [132]. However, as
shown in Fig. 9(d), most of the remaining heat-carrying phonons have intermediate frequency around 0.63 THz. Taking Bi0.5Sb1.5Te3
(bismuth antimony telluride) as an example, ∼74% of the heat is carried by 0.63 THz phonons [132]. These phonons avoid scattering
from boundaries and point defects. However, the dislocations target to scatter the mid-frequency phonons with both dependences of
τD−1∼f and τD−1∼f 3, which is between those for point defect and boundary scattering [133,134]. Thus, dislocations play a
dominant role in phonon scattering effects. Because dislocations always distribute at grain boundaries [132], in order to increase the
density of grain boundaries, nanostructuring has been widely used to realize a high-density of grain boundaries and dislocations
[135–140]. Besides, liquid sintering has also been demonstrated to be useful for further increase the amount of dislocations em-
bedded in grain boundaries [132].

2.4.3. Precipitates
It has been found that the formation of precipitates, especially the nanoprecipitates within the grains, can drastically decrease κl
due to the phonon scatterings caused by these precipitates [8,100,125]. Fig. 9(e) is a TEM image and shows a SnSe matrix with
precipitates [67], and Fig. 9(f) is a HRTEM image of a typical precipitate [67]. The precipitate has a coherent feature and a su-
perstructure according to the fast Fourier transform (FFT) pattern (inset of Fig. 9(f)). Another set of superlattice spots (marked by red
circles) represents a [1 1 1] zone-axis electron diffraction pattern, suggesting that the structure of the precipitate can be identified as Sn
tetragonal phase [67].
An ultralow κl of only ∼0.20 W m−1 K−1 at 773 K was obtained due to the nanoprecipitates obviously decreased κl [67]. Other
experiments [141] demonstrated that rock-salt type SnSe nanoprecipitates coupled with mesoscale matrix grains can lead to further κl
reduction up to ∼0.19 W m−1 K−1 at 850 K. However, it is important to produce appropriate precipitates with suitable size and/or
amount during the fabrication, the misuse of precipitates may cause a decrease of electrical performance, as well as a weak me-
chanical property.

2.5. Influence of pressure

As one of the most important external conditions for the practical applications of thermoelectric materials, the induced external
pressure during their services may distort the crystal lattice and in turn alters the thermoelectric performance [119,142]. Besides, the
study of shear stress as a function of strain under different shear directions is also important for the applications of SnSe thermo-
electric materials.

2.5.1. Crystal structure variation


Fig. 10(a) plots the temperature-pressure contour of electrical resistivity of SnSe, indicating the transition region from semi-
conductor to metallic or semi-metallic region under a pressure [143]. The crystal structure is distorted with increasing the pressure.
To study the influence of pressure on SnSe crystal structures experimentally, the well-crystallized SnSe sample showed that a con-
tinuous transition from α-SnSe (Prototype: GeS) to β’-SnSe (Prototype: TlI) at a high pressure of 10.5 GPa [144]. There were no
indications of any discontinuities in the structural parameters as a function of pressure, suggesting that it was a second-order phase
transition [144]. The β’-SnSe phase was closely related to β-SnSe (Prototype: CrB) at the ambient pressure [144]. Other studies show
that for nanostructured SnSe, such phase transition derived from the change of crystallographic parameters during increasing the

292
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

c d 11.5 a
b
Fractional coordinates

0.8
11.0 c

a, b, c (A)
0.6 xSe o

zSn 10.5
0.4 zSe
xSn 4.0
0.2

3.5
0 3 6 9 12 15 18 0 24 6 8 10 12 14 16 18
Pressure (GPa) Pressure (GPa)
Fig. 10. (a) Temperature-pressure contour plot of resistivity, reproduced with permission [143], Copyright 2016, Royal Society of Chemistry, (b)
ball-and-stick models of SnSe at 5.6 and 8.2 GPa in three perpendicular projections (the yellow spheres represent Se atoms and the grey spheres
represent Sn atoms), (c) pressure-dependent fractional atomic coordinates, and (d) pressure-dependent crystallographic parameters [145].

pressure, and the phase transition occurred between 5.6 and 8.2 GPa [145]. Beyond 8.2 GPa, no significant structural changes were
observed, at least up to 18.0 GPa.
Fig. 10(b) illustrates the projected crystal structures of α-SnSe along different axes in the unit cell at 5.6 and 8.2 GPa [145], in
which the structure is stable at 5.6 GPa. However, at 8.2 GPa, the projections along the b- and c-axes show similar atomic sequences
arranged in a very distorted manner. Along the a-axis, the atoms tend to align along the (0 1 1) plane due to the above mentioned
alignment [145]. Fig. 10(c) shows the change in the fractional coordinates obtained from the Rietveld refinements with increasing
the pressure [145]. Two regions of stability are found (before 4 GPa and after 8 GPa), and the shaded area represents the region of
structural transition. Fig. 10(d) illustrates the change of crystallographic parameters with increasing the pressure [145]. Other studies
[143] showed that SnSe transforms from a semiconducting to semi metallic state under high pressure of 12.6 GPa, followed by a
structural transition from orthorhombic to monoclinic. The average grain size of SnSe decreases when the pressure increases [143].
However, if the pressure continues to increase up to 58 GPa, SnSe exhibits its superconductivity at 4.5 K, and shows a type of CsCl
structure [146].
Compared with the condition at the normal pressure, induced high pressure can result in a phase transition from α-SnSe to β-SnSe,
occurring at a much lower temperature. Because β-SnSe possesses a much higher electrical performance and a much lower κ than α-
SnSe [61], the application of pressure (or compressive strain) may enable SnSe to exhibit improved thermoelectric properties at
temperatures lower than the ∼800 K, which is promising for the practical applications [147]. To verify the influence of pressure on
the electrical performance of SnSe, theoretical calculations using density functional theory (DFT), ab initio molecular dynamics
(AIMD) and Boltzmann transport theory showed that the hydrostatic pressure can reduce the transformation temperature from α-
SnSe to β-SnSe, which contributes to the enhancement in σ/τ and S2σ/τ of SnSe in the middle-temperature range [148], as shown in
Fig. 11(a–c) and (d–f). Here τ is the relaxation time [148].
In terms of κ, the change of crystal structure (and phase transition) influenced by induced high pressure can alter the anhar-
monicity to a great extent [148]. Calculations [148] showed that at 0 GPa, the average γ values along the a-, b- and c-axes are 3.75,
1.9 and 2.6, respectively, indicating a strong anharmonicity. However, when the hydrostatic pressure increased to 4 GPa, the average
γ values along the a-, b- and c-axes became to 1.70, 1.94 and 1.86, respectively, indicating the κ values along a- and c-axes both
increase [148]. Such a larger decrease in the γ values along the inter-layer direction (a-axis) could be related to the higher com-
pressibility along a-axis due to the weak van der Waals force between layers [148]. Fig. 11(g) shows the influence of the induced high
pressure on κ, showed that the high pressure at high temperature can increase κ due to the compressed lattice. Other calculations also
suggested that the induced pressure can significantly enhance the thermoelectric transport properties along all three directions, and

293
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 100 b 800 c 5
a-axis a-axis

)
700

-1
b-axis b-axis

μW cm K S
) 4
-1 -1

c-axis 600 c-axis

-2
10
Sm s

)
-1

-1
500 -SnSe -SnSe

S( VK
0 GPa 3
1 400 a-axis
17

b-axis 2
(10

300 -SnSe -SnSe


c-axis

10
0.1 200

(10
/

-SnSe 1
-SnSe 100
0 GPa 0 GPa

S /
0.01 0 0

2
300 450 600 750 900 300 450 600 750 900 300 450 600 750 900
T (K) T (K) T (K)
d 100 e 800 f 5
a-axis
a-axis

)
a-axis

-1
700 b-axis

μW cm K S
b-axis b-axis 4
)

-2
c-axis
-1 -1

10 c-axis 600 c-axis


Sm s

-SnSe
)

-1
-1

500 3 4 GPa
S( VK

1 4 GPa 400
17

-SnSe 2
(10

300 -SnSe

10
0.1 4 GPa -SnSe

(10
-SnSe -SnSe 200
1
/

100

S /
0.01 0 0

2
300 450 600 750 900 300 450 600 750 900 300 450 600 750 900
T (K) T (K) T (K)
g h
0.35 3.0
a-axis
b-axis 2.5
0.30
Shear Stress (GPa)

c-axis
min (W m-1K-1)

2.0
0.25
0 GPa
(100)/<010>
1.5 (100)/<001>
0.20 (001)/<010>
a-axis 1.0
l

0.15 b-axis 0.5


c-axis
0.10 4 GPa 0.0

200 250 300 350 400 450 500 0.0 0.1 0.2 0.3 0.4
T (K) Shear Strain
Fig. 11. Temperature-dependent electrical transport properties (a) σ/τ, (b) S, (c) S2σ/τ at 0 GPa and (d) σ/τ, (e) S, (f) S2σ/τ at 4 GPa, (g) tem-
perature-dependent κ at both 0 GPa and 4 GPa [183], and (h) calculated shear stress as a function of strain under different shear directions. Here τ is
the relaxation time. Reproduced with permission [149]. Copyright 2017, ACS Publications.

makes SnSe efficient in the medium temperature ranges [142]. Even though, the κ values at 6 GPa are ∼2 times larger than those at 0
GPa, benefited from the huge improvement of electrical performance, the estimated ZT values of p-type SnSe along the b- and c-axis
can reach 2.5 and 1.7 at 6 GPa at 700 K, respectively, and the a-axis shows potential n-type behavior with a ZT value of 1.7 at 600 K
[142].

2.5.2. Shear stress-strain relationship


In practical applications, it is important to choose appropriate pressure values and directions on SnSe thermoelectric materials to
avoid the damage. To study the shear stress-strain relationships of SnSe at 0 K along various slip systems, Fig. 11(h) plots the
calculated shear stress as a function of strain under different shear directions [149]. The (1 0 0)/〈0 1 0〉 and (1 0 0)/〈0 0 1〉 systems
show a similar modulus, and both are lower than that of the (0 0 1)/〈0 1 0〉 system, indicating the (1 0 0)/〈0 1 0〉 and (1 0 0)/
〈0 0 1〉 systems are much weaker than the (0 0 1)/〈0 1 0〉 system in resisting the external deformation [149]. Then the shear stress in
both the (1 0 0)/〈0 1 0〉 and (1 0 0)/〈0 0 1〉 systems gradually increases until reaching the ideal strength. However, in the (0 0 1)/
〈0 1 0〉 system, the shear stress suddenly drops to 0.76 GPa at a strain of 0.07, and then sharply increases to the point of maximum
stress, indicating a significant structural rearrangement at that point. Beyond the maximum stress, the decreased shear stress re-
presents a softening stage of SnSe before the failure. Finally, the shear stress suddenly drops to zero, indicating that the systems can
no longer resist the external deformation, leading to the structural failure. Along the (1 0 0)/〈0 0 1〉 system, a lowest shear strength of

294
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

0.59 GPa is obtained due to the weak van der Waals-like SeeSn bond, suggesting that it is the likely plausible slip surface under
pressure [149]. The maximum shear strength (0.78 GPa) along the (1 0 0)/〈0 1 0〉 slip system is slightly higher than the (0 0 1)/
〈0 1 0〉 system (0.59 GPa). The (0 0 1)/〈0 1 0〉 slip system has a higher ideal strength of 2.81 GPa due to the much stronger covalent
SeeSn bond than the van der Waals-like SeeSn bond in resisting shear deformation [149].

2.5.3. Mechanical properties


For polycrystalline SnSe, the mechanical properties, including compressive strength, bending strength, hardness, fracture
toughness, fracture strength, and thermal shock resistance, all exhibits values that are comparable to those reported for other state-of-
the-art thermoelectric materials. The compressive strength and bending strength of the as-synthesized polycrystalline SnSe were
74.7 MPa and 40.6 MPa [150], respectively. The hardness values measured using Vickers indentation method [151] at a load of
0.98 N with a dwell time of 10 s was found to be 0.27 ± 0.05 GPa. Such a relatively low hardness value of SnSe was attributed to the
brittle nature of SnSe. The fracture toughness (KIC) was found to be 0.76 ± 0.05 MPa m1/2, measured by SEM on the crack lengths of
SnSe [90]. This value is comparable to the other commercial thermoelectric materials, such as PbTe [152] and Bi2Te3 [153]. The
compatibility factor of the as-synthesized SnSe was found to be 0.68 at 773 K [90]. Because compatibility factor for any given
material should be less than 2 to ensure a better power output, the value of 0.68 is also comparable to the other thermoelectric
materials, such as Cu2Se [154] and half-hausler alloys [155]. The thermal shock resistance is another essential parameter during the
actual operation of the device. The temperature gradient between the hot and cold ends of the thermoelectric device legs can lead to
material failure due to the internal mechanical stress caused by the temperature gradients [156–158]. The thermal shock resistance
parameter (RT) of SnSe was found to be ∼252 ± 9 W m−1 [90], which is much higher than those of most state-of-the-art thermo-
electric materials [159]. Therefore, SnSe is a good candidate for the application in thermoelectric devices.

2.6. Oxidation behavior

During the synthesis process of SnSe, it is essential to avoid oxidation occurring on the surface of SnSe to secure a pure phase.
Besides, for the applications of SnSe-based thermoelectric materials, it is also important to avoid oxidation due to the fact that the
oxidized product may seriously affect the properties.
To determine the oxidation of SnSe, X-ray photoelectron spectroscopy (XPS), XRD and DTA analyses have been used. At low
temperatures (<473 K), oxidation of SnSe took place on their surfaces [160]. With increasing the temperature in the range from 523
to 923 K, intermediate tin oxoselenides (presumably SnOSe or SnSeO2) form and subsequently converse to SnO2 [160]. Combined
with analyses from XRD, optical microscopy, scanning electron microscopy (SEM) and energy dispersive spectrometry (EDS), the
oxide thickness has been observed more than 20 mm after oxidization at 873 K for 7 h [161]. Fig. 12(a) shows the XRD pattern of
SnSe surface oxidized under this condition [161], indicating that the surface contains SnO2 and SnSe2. In terms of the oxidation
behavior at room temperature, a combination of aberration-corrected analytical TEM and EDS analysis show that a second thin
amorphous layer can form underneath the outmost SnO2 oxide layer [162].
Fig. 12(b) shows HRTEM image of the surface structure of oxidized SnSe [162], indicating that two different layers formed from
the surface of SnSe after oxidation in air at room temperature for ∼24 h. The outer surface layer marked as L1 has a width of ∼2 nm,
and an inner layer marked as L2 sandwiched between the L1 and the SnSe substrate has a width of ∼4 nm. Fig. 12(c) shows a FFT
pattern of SnSe substrate [162], indicating a typical SnSe crystalline. However, the FFT pattern of L1 and L2 indicates a typical
amorphous structure [162], as shown in Fig. 12(d). To study the composition of L1 and L2, EDS mapping was used. Fig. 12(e)–(h)
show the corresponding EDS mappings of Se, Sn, O, and mixture of Sn and Se, respectively [162]. Clearly, the thin amorphous oxide
layer (L1) was predominantly SnO2, and the second amorphous layer (L2) was Sn1-xSe alloy [162]. To illustrate the proposed for-
mation mechanism of Sn1-xSe layer, Fig. 12(i) shows the schematic free energy versus composition diagram of Sn and Se [162]. The
equiatomic compound (SnSe) has a narrow homogeneity range, and its free energy rises rapidly when the chemical composition
deviates away from the stoichiometric SnSe. During the oxidation process, the consumption of Sn from SnSe to form SnO2 drives the
composition to be Se rich [161]. With increasing the Se concentration, free energy elevated to a none-equilibrium state is higher than
the amorphous phase in the Sn–Se alloy system (the dotted line). This unstable phase can then collapse into the amorphous state,
forming an amorphous layer with an average composition Sn1-xSe [162], and decompose and eventually evolve into crystalline SnSe
and SnSe2 if given sufficient kinetics [162].
The oxidation can seriously damage the thermoelectric performance of SnSe and result in a high κ, contributed by the oxidation
products, mainly SnO2 [68]. Various experiments [163–166] showed that SnO2 possesses a much higher κ than SnSe (about 60 times
of single SnSe). Because the ZT values of SnSe are much higher at high temperature, it is important that SnSe-based thermoelectric
device should prevent to be used in air for long time, and needs to be used under vacuum or with a protective coating such as pure Si
[161]. How to prevent the oxidation of SnSe during the synthesis, measurements, and applications needs urgent attention.

3. Computational advances in SnSe

3.1. Electronic structure

The electronic structures of SnSe-based materials, including the band structure and DOS effective mass, are closely linked to their
thermoelectric performance.

295
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

i
a-SnxSey
Free Energy

SnSe2
SnSe
0 20 40 60 80 100
Sn Atomic Percent of Se Se
Fig. 12. (a) XRD pattern of SnSe surface oxidized at 873 K for 7 h, reproduced with permission [161], Copyright 2016, Elsevier Publishing, (b) Cs-
corrected HRTEM image of the surface structure of oxidized single crystal SnSe, the FFT patterns corresponding to (c) the yellow boxed region and
(d) the blue boxed region. EDS mappings of (e) Se, (f) Sn, (g) O, and (h) mixture of Sn and Se, and (i) schematic free energy versus composition
diagram to illustrate the proposed formation mechanism of α-Sn1-xSe layer. Reproduced with permission [162], Copyright 2017, Springer.

3.1.1. Band structure and effective mass


We first discuss the bandgap (Eg) of SnSe [61]. In 1990s, optical-absorption measurements found an optical bandgap value of
0.923 eV for α-SnSe [167]. With the rapid development of computational science and engineering, theoretical simulations such as
first-principle calculations based on DFT are used to calculate the electronic band structure, as well as the bandgap [168–178].
Table 2 summarizes the recently reported data of the bandgap for SnSe and their calculation methods. As shown in Table 2, different
groups reported different values of bandgap for α-SnSe and β-SnSe [179,180]. The average bandgap of β-SnSe is only roughly half of
that of α-SnSe, indicating there exists an obvious difference between α-SnSe and β-SnSe.
To compare the band structures of α-SnSe and β-SnSe, first-principle calculations based on DFT and many-body perturbation
theory were used and results are shown in Fig. 13(a) and (b) [181]. It is of interest to note that both of them have a very complex
electronic structure with indirect bandgap. The calculated values are 0.83 eV for α-SnSe, and 0.46 eV for β-SnSe, respectively [179].
To illustrate the detailed parameters of these two band structures, Table 3 summarizes the Brillouin-zone positions, energies, and
multiplicities of the band extrema of both α-SnSe and β-SnSe, where k1, k2, and k3 are the coordinates in the unit cell [181]. All
energies are referenced to the valence band maximum (VBM) of each phase. For α-SnSe, the position of the VBM is at (0.00, 0.35,
0.00) along the Γ–Y direction of the first Brillouin zone, and a local valence-band maximum (VBM1) at (0.00, 0.42, 0.00) lies within
1 meV lower in energy than the VBM. The conduction-band minimum (CBM) is located at (0.33, 0.00, 0.00) along the Γ–X direction.
For β-SnSe, both the VBM and CBM are located at (0.34, 0.50, 0.00) along the X–A direction. In addition, a VBM1 at (0.00, 0.20, 0.39)

296
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Table 2
A summary of bandgap data of SnSe and calculation methods.
Structure (Calculated) Bandgap (eV) (Calculation) Method Year Ref.

α-SnSe, β-SnSe 0.61, 0.39 First-principles density 2014 [61]


functional theory
α-SnSe, β-SnSe 0.829, 0.464 Density functional and many-body perturbation theory 2015 [181]
α-SnSe, β-SnSe 0.86, 0.355 Full-potential Korringa-Kohn-Rostoker 2015 [179]
α-SnSe 0.44 First-principles density 2015 [188]
functional theory
α-SnSe 0.60 First-principles density 2015 [79]
functional theory
α-SnSe 0.643 First-principles density 2012 [267]
functional theory
α-SnSe 0.63 First-principles density 2017 [197]
functional theory
α-SnSe 0.93 First-principles density 2017 [384]
functional theory
α-SnSe 0.65 First-principles density 2017 [317]
functional theory with Vienna Ab Initio Simulation Package
α-SnSe 0.78 Ab initio many-body perturbation theory 2015 [110]
α-SnSe 0.86 Vienna Ab Initio Simulation Package 2015 [117]
α-SnSe 0.60 Vienna Ab Initio Simulation Package 2015 [183]
α-SnSe 0.80 Vienna Ab Initio Simulation Package 2016 [90]
α-SnSe 0.935 Full-potential linearized augmented plane waves 2015 [204]
α-SnSe 0.69 Full-potential linearized augmented plane waves 2015 [194]
α-SnSe – Full-potential linearized augmented plane waves 2015 [177]
α-SnSe 0.923 Optical-absorption measurements 1981 [167]
α-SnSe 0.86 mBJ method 2015 [196]

along the Γ –H direction and a local conduction-band minimum (CBM1) at (0.00, 0.54, 0.08) along the X–H1 direction can be found.
Both phases exhibit multiple local band extrema near their band edges that need to be considered when evaluating the thermoelectric
properties for SnSe [181]. Besides, Table 3 lists the reported effective mass (m∗) along different directions for both α-SnSe and β-SnSe
[181]. Clearly, m∗ at each extremum is highly anisotropic: ma∗ has a larger value along the a-axis than that of the in-plane (b-axis
(mb∗) and c-axis (mc∗)). This is due to the two-dimensional nature of the SnSe structure, which favors electron transport within the
atomic layers [181].

3.1.2. Density of states and Fermi energy


As a typical crystalline material, SnSe exhibits multiple extrema in its valence or conduction bands. The increase of the local DOS
near the Fermi level can greatly enhance S, and in turn enhance S2σ [68,180,181]. S can be expressed by the Mott equation [79,182]:

π 2kB2 T ⎧ dn (E ) dμ (E ) ⎫
S= +
3q ⎨ ⎩ ndE μdE ⎬ ⎭E = EF (5)

with
n (E ) = g (E ) f (E ) (6)

where kB is the Boltzmann constant, f(E) is the Fermi function, q is the carrier charge, n(E) is the carrier concentration, μ(E) is the
carrier mobility, and g(E) is the DOS. Thus, S depends on the energy derivative of the DOS. For a given n, it is well known that a high
m∗ is beneficial for a high S. To study the contribution of each atom to the DOS in SnSe, first-principles and semiclassical Boltzmann
theory were used. Fig. 13(c) and (d) show typical calculated DOS of α-SnSe and β-SnSe, respectively [180]. A finite gap between
valence and conduction bands can be seen for both α-SnSe and β-SnSe, and the former one is 0.3 eV larger than the latter one. The
valence band comprises of peaks at ∼−7.5 eV and between 0 and −5 eV. The top of the valence band has major contributions from
Se-p and Sn-s while the bottom of the conduction band arises from Sn-p and Se-p states. Conduction band mainly comprises of Sn-p
and Se-p states. Meanwhile, it is obvious that the DOS of α-SnSe is higher than that of β-SnSe, indicating that α-SnSe has higher S than
β-SnSe. The decrease of the DOS from α-SnSe to β-SnSe is mainly due to the decrease of Sn p-states and Se p-states, which is the main
reason for S decrease in β-SnSe [180].
Fig. 13(e) and (f) show the calculated Fermi energy [180] for α-SnSe and β-SnSe as a function of temperature with a hole
concentration of 6 × 1017 cm−3. The Fermi energy lies within the bandgap for both phases at 300 K, with 0.1 eV above the VBM for
α-SnSe and 0.05 eV above the VBM for β-SnSe. With increasing the temperature, electrons are thermally excited from the valence
band to the conduction band, and the Fermi energy shifts towards the middle of the bandgap. The Fermi energy of β-SnSe stabilizes
near the middle of the gap for temperatures above 700 K, indicating a comparable concentration of free electrons and holes and is
consistent with the sign reversal of S at the high temperatures [61,81,180]. It is of interest to note that the experimental peak S value
(marked as Smax) of SnSe single crystal is different from the calculated values [61]. The experimental Smax of SnSe single crystal is
found between 500 K and 600 K (marked as T∗) due to the bipolar transport [61], while the calculations showed that the onset of the

297
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

e f
0.8
Fermi Energy (eV)
Fermi Energy (eV)

0.4
0.6 β-SnSe
α-SnSe
0.4 0.2

0.2
0.0
0.0

300 450 600 750 900 300 450 600 750 900
T (K) T (K)
Fig. 13. Calculated SnSe electronic band structures of (a) α-SnSe and (b) β-SnSe, reproduced with permission [181], Copyright 2015, AIP Pub-
lishing, the total and partial DOS of (c) α-SnSe and (d) β-SnSe, reproduced with permission [180], Copyright 2016, Springer Nature, and the
temperature-dependent Fermi energy of (e) α-SnSe and (f) β-SnSe [181].

Table 3
A summary of calculated values of the positions, energies of the conduction and valence band extrema, and effective masses along the a, b, and c-
axes [181].
Multiplicity (k1, k2, k3) E (eV) ma* (me) mb* (me) mc* (me)

α-SnSe, VBM 2 (0.00, 0.35, 0.00) 0.000 0.74 0.31 0.16


α-SnSe, VBM1 2 (0.00, 0.42, 0.00) 0.000 0.90 0.12 0.15
α-SnSe, CBM 2 (0.33, 0.00, 0.00) 0.829 2.40 0.11 0.15
β-SnSe, VBM 2 (0.34, 0.50, 0.00) 0.000 0.34 0.04 0.09
β-SnSe, VBM1 2 (0.00, 0.20, 0.39) −0.031 0.77 0.12 0.05
β-SnSe, CBM 2 (0.34, 0.50, 0.00) 0.464 3.07 0.04 0.10
β-SnSe, CBM1 2 (0.00, 0.54, 0.08) 0.534 0.06 0.82 1.52

bipolar transport should occur at a higher temperature (T∗ = ∼700 K). This is because the calculations have not included the effect of
the temperature on the calculated band structure, which decreases the bandgap with increasing the temperature and reduces the
temperature onset of the bipolar transport [181].

298
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

d 0.8

Volume (A3)
o
240

0.6

Energy gap (eV)


230

220 0.4

210
0.2
Strain (%)
200
-3 -2 -1 0 1 2 3

Fig. 14. (a) Band structures of the unstrained SnSe. Inset: Brillouin zone of the orthorhombic SnSe with labels of high symmetry k points, Colors
represent p orbitals, the electronic band structure of SnSe in (b) −3% compressive strain and (c) +3% tensile strain, (d) dependence of volume and
bandgap with different strain level from −3% to +3% [183].

3.2. Influence of the strain

As discussed in Section 2.4, strain can be induced by the point defects or dislocations in SnSe materials. In addition, doping in
SnSe can cause strain through the substitution of dopant atoms with different sizes.

3.2.1. Band structure modification


DFT calculations was used to study the influence of strain on electronic structure of SnSe [183]. Fig. 14(a) shows a typical
electronic structure of the unstrained SnSe [183] with an inset indicating the Brillouin zone of α-SnSe. The unstrained SnSe exhibits
an indirect bandgap of 0.60 eV. As can be seen, threefold degeneracy of p orbitals of both Sn and Se are lifted due to the orthorhombic
structure [183]. The VBM is mainly of Se 4py state with small contribution from Se 4px state, while the CBM is dominated by 5px and
5pz states. The topmost valence band is quite dispersive along the Y–Γ and Z–Γ directions with the calculated hole effective mass
(mh∗), which is 0.16 me at the VBM, while it is flat along the Γ–X direction due to the layered structure [183]. The lowest conduction
band along the Z–Γ direction is also dispersive around the CBM with the electron effective mass me∗ = 0.137 me [183]. Compared to
the CBM, conduction bands along the Γ–X direction are very dispersive, but located at rather higher energy. Therefore, it gives less
effects to the thermoelectric properties in the unstrained SnSe [183].
To study the strain effect on band structure, the strains from −3% (compressive strain) to +3% (tensile strain) were applied in
the bc-plane. Fig. 14(b) and (c) show the band structures applied by −3% and +3% strains, respectively [183]. When strains were
applied, the band structure and bandgap values were significantly affected, and the CBM and VBM were also shifted with change in
bandgap [183]. For +3% strain, the VBM occurs at (0.00 0.33 0.00), while for −3% strain, it occurs at (0.00 0.43 0.00). The −3%
strain shifts the VBM toward Y in Γ−Y, while the same orbital characters are retained regardless of −3% or +3% strains. Besides,
the change in conduction band is quite different as these results from the level reversal between Sn px and pz orbitals. For the −3%
strain, the CBM occurs closer to Γ with dominant Sn px character, while for the +3% strain, it occurs close to Y with dominant Sn pz
contribution. Meanwhile, for the +3% strain, many valleys in the CBM are pushed upward, while for the −3% strain, those valleys
moves downwards much closer to the VBM. As such, the strain on the bc-plane affects orbitals along the a-axis, which is well
manifested with shifts of Sn px in conduction bands. However, the changes in valence states are not as dramatic as conduction bands
[183]. Fig. 14(d) shows volumes and bandgap values for different strains. It is obvious that tensile strains increase volumes and
bandgaps, whereas compressive strains reduce both.

3.2.2. Thermopower optimization


S exhibits a strong dependence on the applied strain. Fig. 15(a)–(c) show calculated effect of strain on S of SnSe in along a-, b- and

299
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 15. Effects of strain on S along (a) a-axis, (b) b-axis, and (c) c-axis [183], (d) cell volume varying with alloy composition, (e) κl with the
temperature for 5 mol% dopant amount, and (f) κl with different alloyed dopants (S, Ge, Te, Sr, Ba). Reproduced with permission [143], Copyright
2016, Royal Society of Chemistry.

c-axes [183]. All calculations are performed with n = 6 × 1017 cm−3. S anisotropy is well exhibited derived from the orthorhombic
structure. It is of interest to note that the temperature of Smax (marked as T∗) increases from 525 K for 0% strain to 700 K for +3%
strain, whereas T∗ reduces to 350 K for −3% strain. The strain-dependent T∗ is attributed to the change of bandgap under the strain
[183]. Strains can significantly affect the calculated S. For the −3% strain, S changes significantly. Sb and Sc are strongly reduced
compared to other cases. The reduction of Sa is much more drastic, implying the system becoming n-type. The change of S at −3%
stems from the change in band structure. In lattices without strain or with the +3% strain, the CBM occurs between Z and Γ, while for
the −3% strain, the CBM appears nearly at Γ. Furthermore, due to the orbital reversal, conduction bands become highly dispersive
for the −3% strain. The reduction of S slows down by the applied +3% strain at the high temperature, which means that applying
appropriate strain can increase S at the high temperature. Compressive strain can make thermal excitation much easier, where the
bipolar transport (a coexistence of n- and p-type carriers) can be realized [183].
As discussed above, doping with other elements may cause strain due to the substitution of dopant atoms with different atom size
from Sn or Se atom. To investigate the lattice change caused by induced strain, Fig. 15(d) shows that the unit cell volume varies with
alloyed compositional species (S, Ge, Te, Sr, and Ba) with a solubility of >2 mol% [127]. The unit cell volume for pure SnSe (black) is
given for comparison. It demonstrates a linear dependence on the alloyed composition, following the classical Vegard’s law [184].
Fig. 15(e) shows the plots of κl as a function of temperature for different dopants with 5 mol% doping amount [127]. It is clearly seen
that the Debye–Callaway model [185,186] generates excellent fits for the temperature-dependent κl, indicating that the scattering
physics is mainly determined by a combination of Umklapp, boundary, and point defect scattering. Fig. 15(f) shows the plots of κl
with changing the doping amount of different dopants [127], which fits the Abeles model [187]. With increasing the doping amount,
the density of point defects increases, resulting in a decreased κl. Meanwhile, with increasing the atom size of dopant, κl decreases,
indicating that the selection of heavy element is a suitable way to further decrease κl.

3.3. Energy band engineering

As discussed above, doping is one of effective routes to further decrease κ. In addition, the induced dopants can also produce new
chemical or ionic bonding, which in turn alter the band structures and bandgaps of the materials [188], and improve the thermo-
electric properties to a great extent [179].

3.3.1. n/p-type doping


To elucidate how dopants modify the band structure and thermoelectric properties of SnSe, theoretical investigations based on the
first-principles calculations and the Boltzmann transport theory were used [189,190]. Fig. 16(a) and (b) show the calculated band
structure of n-type Bi0.028Sn0.972Se (green solid line) and p-type Ag0.028Sn0.972Se (blue solid line) at 300 K, and these compositions
correspond to the n of 4.79 × 1020 cm−3 [188]. The entire bands are rigidly shifted along the y-axis, so that the chemical potential at

300
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 16. Calculated band structures of (a) n-type Bi0.028Sn0.972Se (green solid line) and (b) p-type Ag0.028Sn0.972Se (blue solid line) by super cell
model, in which the band structure of the un-doped SnSe (red dashed line) is also shown. Reproduced with permission [188]. Copyright 2015, Royal
Society of Chemistry. Calculated band structures of (c) undoped and (d) 3% Na-doped α-SnSe by super cell model. Reproduced with permission
[79]. Copyright 2015, Royal Society of Chemistry.

300 K is located at zero for each system. The band structure of the un-doped SnSe (red dashed line) [188] is also shown in both
panels. Considerable changes in band structures have been observed around the CBM of Bi0.028Sn0.972Se and around the VBM of
Ag0.028Sn0.972Se, respectively [188]. Besides, the bandgap slightly narrows through the doping process. Other DFT calculations [79]
showed the similar results. Fig. 16(c) and (d) show the calculated band structures for undoped and 3% Na-doped α-SnSe using the
super cell model, respectively [79]. From which, doping with 3% Na can modify the band structure by shifting the Fermi level, as well
as flattening the top of the valence band, which narrows the bandgap [79].
Fig. 17(a) shows the DOS near the Fermi level [79], which is derived from the s- and p-orbitals of Sn and especially from the p-
orbitals of Se. As can be seen, the contribution of the Na dopants to the DOS is almost negligible. Fig. 17(b) shows the calculated DOS
with different Na doping concentrations (namely 1%, 2%, 3%, and 4%), from which altered Fermi levels can be seen [79]. In
particular, the Fermi level decreases with increasing the Na content, and the top of the VB remains at ∼3.5 eV. Fig. 17(c) depicts this
result by the red solid line [79], where the top bands refer to VB1 or VB2 and the sub-bands refer to VB3 or VB4. The dopants change
the profile of the original DOS by increasing the number of carrier pockets near the Fermi level. Fig. 17(d) shows the Fermi level shifts
with increasing the temperature [79], deriving from n decrease during increasing the temperature.

3.3.2. Carrier concentration tuning


Doping can be used to effectively tune n, and in turn alter the electrical transport. Eqs. (5) and (6) indicate that S depends on the
energy derivative of the DOS (g(E)), as well as n or n(E) since n(E) = g(E)f(E). Both the calculations [181] and experiments [79]
demonstrated that with increasing n, the onset of bipolar transport should occur at the higher temperatures, which is beneficial to
enhance the thermoelectric performance at the higher temperature. However, Smax may decrease to some extent. More frequently
used relationship between S and n can be described as [191–193]:

8π 2kB2 ∗ π 2/3
S= 2
m T⎛ ⎞
3eh ⎝ 3n ⎠ (7)

and
σ = neμ (8)

where h is the Planck's constant. n can increase via doping of SnSe, resulting in an increase of σ [113]. Experiments demonstrated that
by doping with Na, n increased from ∼3 × 1017 cm−3 for undoped SnSe [61] to ∼3.2 × 1019 cm−3 for 1% Na-doped SnSe, and
∼5.2 × 1019 cm−3 for 2% Na-doped SnSe [79]. Although S of Na-doped SnSe is distinctly smaller than that of undoped SnSe, much
higher σ leads to larger S2σ compared to the undoped SnSe [61,79]. Thus, designing the suitable type and amount of dopant(s) plays a

301
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

c Forbidden zone d
3.8 3.8 Forbidden zone
Energy (eV)

Energy (eV)
3.6 Top bands Top bands
3.6
EC EC
DoS increase

DoS increase
3.4 Esub
3.4
Esub

3.2 3.2 Sub-bands


Sub-bands
0 1 2 3 4 300 400 500 600 700 800
Doping concentration (%) T (K)
Fig. 17. Calculated (a) partial DOS for 3% Na-doped α-SnSe, (b) total DOS of Sn1-xNaxSe (x = 0, 0.01, 0.02, 0.03, 0.04), and Fermi level as a
function of the doping concentration (c) and the temperature (d). Reproduced with permission [79]. Copyright 2015, Royal Society of Chemistry.

dominant role in further improving the thermoelectric performance of SnSe.

3.4. Summary of calculation-based work

Calculation materials science is essential for the verification of current experimental results, as well as the prediction of potential
properties of materials and device performances. For this reason, extensive calculations have been performed in the SnSe materials
system.

3.4.1. Verification of experimental results


Since SnSe has been reported as a promising thermoelectric system, many calculations were performed
[90,109,110,113,117,179–181,188,189,194–200]. To verify the experimental results of single crystal SnSe [61], by using the
equation of [188]:

n= ∫ D (ε ) exp [(ε−μ 1)/kB T ] + 1 dε (9)


i

17 −3
where D(ε) is DOS, n = 6 × 10 cm was obtained, from which the calculation successfully reproduced the experimental data
[188]. Here the chemical potential μi is the parameter to adjust n. When choosing n = 3.3 × 1017 cm−3, the calculated electrical
transport properties (such as σ and S) were all in good agreement with experimental results, and resulting in a very high ZT value of
2.1 at 923 K [113], which was a little lower than experimental value (ZT = 2.6 at 923 K) [61].
For polycrystalline SnSe, calculations based on the DFT and Boltzmann transport equations showed a reasonable agreement with
the theoretically derived band gap and experimentally measured temperature-dependent σ and S [90]. These calculation results
provide the evidence of the authenticity of experiment results.

3.4.2. Theoretical prediction


As discussed in Section 3.3, decoupling σ and S via tuning n through band engineering is a key strategy to improve ZT [48,201].
Extensive experiments show that the highest ZT can be approached for the single crystal SnSe along the b-axis when n can be
optimized within the range of 1019–1020 cm−3 [79]. To predict the optimized n, calculations based on full-potential linearized
augmented plane waves (FLAPW) method and the semiclassical Boltzmann theory were used [202,203]. Fig. 18 shows the calculated
σ, S, S2σ, and ZT for both n-type and p-type SnSe at different n values at 770 K [204]. Fig. 18(a) shows the plots of σ as a function of n,
in which along the a-axis, n-type doping has much larger σ than that of p-type doping, but along the b-axis, σ of n-type doping is lower

302
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

600
a 10 a-axis, p-type b a-axis, p-type
b-axis, p-type b-axis, p-type
400
a-axis, n-type
8

σ (104 S cm-1)
b-axis, n-type 200

S (μV K-1)
6
0
4 -200
2 -400 a-axis, n-type
b-axis, n-type
0 -600
19 20 21
19 20 21
10 10 10 10 10 10
n (cm-3) n (cm-3)
c 25 a-axis, p-type d 4
a-axis, p-type
b-axis, p-type b-axis, p-type
S2σ (μW cm-1K-2)

a-axis, n-type a-axis, n-type


20 b-axis, n-type b-axis, n-type
3
15

ZT
2
10
1
5

0 19 20 21
0 19 20 21
10 10 10 10 10 10
n (cm-3) n (cm-3)
Fig. 18. Calculated n-dependent σ, (b) S, (c) S2σ, and (d) ZT at 750 K. Reproduced with permission [204]. Copyright 2015, Elsevier Publishing.

than that of p-type. Fig. 18(b) suggests that along the different axes, S are roughly the same for both n-type and p-type doping.
Therefore, S2σ of n-type doping are much greater than that of the p-type doping along the a-axis, and S2σ of n-type doping are lower
than that of the p-type doping along the b-axis, as shown in Fig. 18(c). The maximum S2σ of n-type is nearly 10 times than that of p-
type along the a-axis. In order to estimate ZT, the experimental κ values were selected [61]. Fig. 18(d) shows that for p-type SnSe, the
ZT value is increased to 2.2 along the b- axis compared with the experimental value of 1.5 at 750 K when n increased to 3.2 × 1019
cm−3 [61]. Besides, the roughly estimated maximum ZT value is ∼3.1 for n-type SnSe along the a-axis at 750 K, corresponding to
n = 2.8 × 1019 cm−3. Although the estimated peak ZT = ∼3.1 for n-type SnSe may not be very accurate due to not directly deduce
relaxation time from experimental results [204], it does suggest the potential to further enhance the thermoelectric performance of
SnSe.
The analysis of the doping effect by rigid-band approximation shows that the enhancement of the ZT value can be achieved by
doping [188], particularly when n is in the range of 5 × 1019–5 × 1020 cm−3 for both p- and n-type doping. A maximum ZT of ∼0.6
at 300 K and ∼1.8 at 450 K can be predicted [188]. Calculations based on the semiclassical model incorporating nonparabolicity, the
two-band Kane model, the Hall factor, and the Debye–Callaway model [113] demonstrated that at n = ∼3 × 1019 cm−3, peak ZT
values can be predicted to be ∼0.5, ∼1.5 and ∼2.3 at temperatures of 300, 600 and 900 K, respectively. In fact, ZT = ∼1.5 at 600 K
can be very useful in some practical applications. Calculations based on DFT demonstrated that for p-type SnSe at 750 K, when
n = ∼6 × 1019 cm−3, peak ZT values can be predicted to be ∼1.0, ∼0.6 and ∼0.5 along b-, c- and a-axes, respectively; while for n-
type SnSe at 750 K, when n = ∼4 × 1019 cm−3, peak ZT values can be predicted to be ∼2.6, ∼1.3 and ∼1.0 along a-, c- and b-axes,
respectively [110]. Calculations based on DFT with the full-potential linearized augmented plane-wave (LAPW) method also de-
monstrated that for p-type single crystal SnSe at 675 K, the peak ZT values along a-, b- and c-axes were ∼1.0, ∼2.6 and ∼1.7 at n of
∼6.2 × 1019, ∼3.6 × 1019 and ∼4.1 × 1019 cm−3, respectively [196]. The highest ZT value of polycrystalline SnSe at the same
temperature was predicted to be 1.86 for n = ∼4.2 × 1019 cm−3 for p-type SnSe [196]. Therefore, tuning n is very important for
securing high ZT, and it possesses full potential for both p-type and n-type doping for SnSe-based thermoelectric materials.

3.4.3. Crystal design


Apart from α-SnSe and β-SnSe, another cubic structured SnSe phase with a lattice parameter a0 = 11.9702 Å and space group of
Fm3m (marked as π-SnSe) was also identified under the condition of hydrothermally synthesized nanoparticles [141,205]. Fig. 19(a)
shows the structure of π-SnSe with 64 atoms per unit cell [206]. Fig. 19(b) shows the calculated band structure of π-SnSe using mBJ
potential and Boltzmann transport theory [206]. As can be seen, π-SnSe has a calculated bandgap of 0.97 eV. Fig. 19(c) shows S, S2σ
and ZT as a function of temperature for π-SnSe [206]. π-SnSe exhibited high S values at low temperature, and the values decreased
from 2292 to 444.3 μV K−1 with increasing the temperature from 200 to 900 K, resulting in a reduction in the ZT values [206]. It is
noticeable that π-SnSe exhibits a higher value of ZT (∼1.0) at relatively low temperature (200 K) when compared to the

303
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

c 2400 60 1.00

S2σ (×1010 W m-1K-2s)


50 0.98

2000 40 0.96

ZT
30 0.94

S (μV K-1)
20 0.92
1600 10 0.90

0 0.88
300 600 900
1200 T (K)
S
800 S2σ
400 ZT

300 600 900


T (K)

f 2.0 single crystal, b-axis


single crystal, c-axis
single crystal, a-axis
1.6
nanotube

κl (W m-1K-1)
1.2

0.8

0.4

300 400 500 600 700 800


T (K)
Fig. 19. (a) Schematic crystal structure of π-SnSe and its atomic position having 64 atoms per unit cell, (b) electronic band structures of π-SnSe
along Σ using mBJ potential, (c) temperature-dependent S, S2σ and ZT of π-SnSe, reproduced with permission [206], Copyright 2017, Elsevier
Publishing, (d) top view of a SnSe nanotube, (e) Electronic band structures of the SnSe nanotube, and (f) Temperature-dependent κl of the SnSe
nanotube compared with that of single crystal SnSe. Reproduced with permission [208]. Copyright 2017, Royal Society of Chemistry.

orthorhombic SnSe [206]. Other computational calculations based on first-principles pseudopotential method showed that the
bandgap of π-SnSe decreased from ∼1.0 eV to ∼0 eV when external hydrostatic pressure was applied and increased from 0 GPa to 40
GPa [207]. These results indicate that π-SnSe has a wide range of applications for pressure based thermoelectric materials.
Fig. 19(d) shows a “star-like” SnSe nanotube with a cylindrical nanostructure, which was designed via first-principles simulations
[208]. Fig. 19(e) shows the calculated band structure of SnSe nanotube with a bandgap of 1.55 eV [208]. Fig. 19(f) shows κl of SnSe
nanotube compared with its single crystal counterpart [208], an ultralow κl of 0.18 W m−1 K−1 at 750 K was obtained. Besides, a
large S of 900 μV K−1 for p-type nanotube and −700 μV K−1 for n-type nanotube were obtained both at 750 K. When doping level
was tuned to around 1020–1021 cm−3, S2σ of the p-type SnSe nanotube reached its maximum value of 235 μW cm−1 K−2, and an
exceptional high ZT value of 3.5–4.6 under the optimal conditions can be calculated [208]. For SnSe nanoribbons [209,210], by using
DFT coupled with semi-classical Boltzmann theory, a relatively high S of ∼1720 μV K−1 and low κ of 0.2 W m−1 K−1 were calculated
for the armchair nanoribbon of 6 Å width [210]. Meanwhile, a large relaxation time τ of 55.57 fs and a small m∗ of 0.11 me, both for
carriers of holes, were obtained for the armchair nanoribbon of 47 Å width [210]. The simulation results showed that SnSe nanor-
ibbons may provide thermoelectric performance that is similar to that of the monolayer and low-temperature bulk of SnSe.

4. Single crystal SnSe

4.1. Crystal growth

To date, there are four main methods to fabricate single crystal SnSe: Bridgman method (including Bridgman-Stockbarger
method) [211–213], direct vapor transport method [214–216], temperature gradient growth method [52,217,218], and hydro-
thermal method (including solvothermal method) [219–222]. Table 4 summarizes the fabrication methods of single crystal SnSe
reported in recent years, where the Bridgman method is abbreviated as B, the Bridgman-Stockbarger method as BS, the direct vapor
transport method as DVT, the hydrothermal method as HT, solvothermal method as ST, and the temperature gradient growth method
[223] as TGG. Within these fabrication methods, HT and ST methods have been frequently applied for the fabrication of small-scaled
single crystal SnSe, and these small crystals are often used as precursors for sintering bulk thermoelectric materials [81,141,224].

4.1.1. Bridgman method


Both the B method (including the BS method) and the DVT method have a similar pre-procedure to fabricate SnSe by melting
[225–227] or solid-state reactions [228–230]. The precursors are high-purity Sn and Se powders or shots (or ingots) with a purity

304
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Table 4
A comprehensive summary on the synthesis of single crystal SnSe. Here the Bridgman method is abbreviated as B, the Bridgman-Stockbarger method
as BS, the direct vapor transport method as DVT, the hydro-thermal method as HT, solve-thermal method as ST, and the temperature gradient
growth method as TGG.
Product Precursors Atmosphere Pressure (Torr) Temperature (K) Heating Keeping Crystal growth Year Ref.
time (h) time (h) method

Sn0.985Na0.015Se 99.999% Sn, Se Vacuum 10−4 1223 10 6 B 2015 [78]


and Na
SnSe 99.999% Sn and Se Vacuum 10−4 1223 10 6 B 2014 [61]
SnSe – – – – – – B 1992 [416]
SnSe – – – – – – B 1989 [417]
SnSe – – – – – – B 1986 [418]
SnSe 99.999% Sn and Se vacuum – 1223 – 20 Modified B 2017 [246]
Sn1-xRxSe (R = Na, – argon ∼3.7 × 10−6 1273 12 10 Modified B 2015 [79]
Ag)
SnSe 99.999% Sn and Se Vacuum – 1223 – 20 Modified B 2015 [115]
Sn0.985Na0.015Se 99.999% Sn, Se Vacuum ∼10−4 1223 10 6 Vertical B 2017 [237]
and Na
Sn0.97Ag0.03Se – vacuum ∼7.5 × 10−6 1183 – 20 Horizontal B 2017 [242]
SnSe – Vacuum ∼7.5 × 10−6 1223 ∼0.8 – Horizontal B 2017 [419]
SnSe 99.999% Sn and Se Vacuum 10−5 1183 – 24 BS 2009 [420]
SnSe 99.99% Sn and Se vacuum 5 × 10−5 1234 – 48 BS 1989 [80]
SnSe 99.99% Sn and Se Vacuum 10−5 1073 ∼13 72 DVT 2009 [216]
SnSe 99.99% Sn and Se Vacuum ∼7.5 × 10−6 1053 – 170 DVT 2007 [92]
SnSe – – – – – – DVT 2000 [231]
SnSe 99.99% Sn and Se Vacuum 10−5 1053 ∼75 168 DVT 1994 [215]
SnSe – – – – – – DVT 1981 [167]
SnSe – Vacuum 10−6 1173 – 4 DVT 1977 [421]
SnSe 99.95% Se and EDTA – 453 – 168 ST 2000 [224]
SnCl2
Sn1-xBixSe 99.999% Sn, Se Vacuum 10−4 1203 90 10 TGG 2016 [52]
and Bi
SnSe1-xSx – Vacuum – – – – TGG 2017 [217]
SnSe1-xSx 99.8% Sn, 99.9% Vacuum <10−4 1233 – 33.3 TGG 2017 [218]
Se and 99% S

of > 99.99% to prevent impurity contaminations in the final fabricated products. For this reason, the transporting agents are needed
[231]. There are mainly five steps for fabricating SnSe. First, the precursors and transporting agents are sealed under high vacuum
(from 10−4 ∼ 10−7 Torr) in a quartz ampoule, then melted or solid state reacted in a furnace with an appropriate temperature and an
appropriate heating speed. For the melting route, the targeted temperature can be set above the melting point of SnSe (1134 K), while
for solid-state reaction, the targeted temperature should be set just below the melting point. The heating speed should not be set too
high to make sure that the Sn and Se are adequately reacted to form SnSe. When the furnace temperature reaches to the targeted
temperature, keep this temperature for a long time (from 12 h to several days) to ensure a full reaction. When the reaction is fully
proceeded, slowly cool the furnace to room temperature. The obtained ingots are crushed into powders and prepared for the single
crystal growth. As discussed in Section 2.6, it should be note that the oxidization is better to be avoided during the fabrication of
SnSe.
When using the B method (including the BS method), a seed SnSe crystal is first located at bottom, then the SnSe powders or ingot
are heated above the melting point and slowly cooled down at the seed SnSe crystal. The single crystal SnSe grows on the seed along
the same crystallographic orientation of the seed material. The growth process is always carried out in a vertical or horizontal
geometry, which is called vertical Bridgman method and horizontal Bridgman method. The temperature controllers are needed to
control the crystal growth process. There is some difference between the B method and advanced Stockbarger method [213]. Both
methods need a moving crucible and a temperature gradient, but the Bridgman route uses the relatively uncontrolled gradient
produced at the exit of the furnace; and the Stockbarger route applies a baffle (or shelf) to separate two coupled furnaces with
temperatures above and below the freezing point. After the improvement by Stockbarger, the BS method possesses better control over
the temperature gradient at the melt or crystal interface [213].

4.1.2. Direct vapor transport method


In the DVT method [214–216], when the homogeneous polycrystalline SnSe powders are obtained from melting or solid-state
reaction, the SnSe powders are then placed in a charged ampoule, and the ampoule is evacuated and sealed at a high vacuum of 10−3
Pa and placed in a dual zone horizontal furnace with a temperature gradient. The charge end (hot zone) is maintained at a higher
temperature (such as 1063 K) and the tapered growth end (cold zone) at the lower temperature (such as 1013–943 K). The tem-
perature is increased at a rate of 10–30 K per hour from room temperature (300 K). When the targeted temperature is reached, the
ampoule is left in the furnace for a long time (such as 240 h, much longer than the B method). After that, the cooling is carried out at a
low rate (such as 5 K per hour) to room temperature. At last, the furnace is switched off and the ampoule is carefully taken out from

305
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

the furnace. The ampoule is finally broken to obtain the single crystals.
It is obvious that single crystal SnSe need rigid conditions to grow. Besides, compared with the synthesis of polycrystalline SnSe
materials, the efficiency of single crystal growth is low, and the cost is high. Further improvement of the growth technique of single
crystals needs to be addressed.

4.1.3. Temperature gradient growth method


For TGG route, the temperature gradient plays a critical role in the growth of single crystal SnSe. Similar to B and DVT methods,
high-purity Sn and Se powders were used as precursors for the crystal growth [52,217,218]. It was better to choose these precursors
with particle sizes <200 meshes to optimize the surface area, and thereby enhance the reaction kinetics [223]. Sn and Se powders
were first loaded into thick wall quartz ampoules. Then, the ampoules were evacuated and sealed. Sometimes another quartz tube
was sealed to protect the sample and ampoule. The powders in ampoule were mixed and loaded into a vertical furnace. The tem-
perature was slowly increased from room temperature to above the melting point of SnSe, and maintained at this temperature for an
appropriate time (such as 10 h [52]). Finally, the temperature was slowly cooled at an appropriate rate to below the melting point of
SnSe (such as 973 K [52]), and then rapidly decreased to room temperature.

4.1.4. Hydro/solvothermal method


HT and ST syntheses [219–222] are also effective methods to obtain SnSe products, such as single crystal plates [81,91,232] and
nanobelts [233]. In HT synthesis, single crystal SnSe is synthesized in hot water under high pressure. The apparatus for taking HT
synthesis is an autoclave (also called steel pressure vessel), in which a nutrient is supplied along with water. The autoclaves are put
into an oven for chemical reactions under high temperature. With increasing the pressure and temperature, the water ions can rapidly
increase. Under a high pressure (15–20 GPa) and high temperature, the density of water can reach 1.7–1.9 g cm−3. In this condition,
water can be completely dissociated into OH− and H3O+ (behaving like a molten salt). Besides, the water viscosity decreases with
increasing the temperature. In this situation, the mobility of ions and molecules is much higher than under normal condition. Many
chemical reactions can be carried out in this environment. Meanwhile, during the chemical reaction, water is used as a solvent, and
the dielectric constant is one of the very important properties of water. With increasing the temperature and pressure, the dielectric
constant of water decreases [234]. Therefore, temperature plays a dominant role in the dielectric constant of water. In terms of ST
synthesis, it is similar to the HT synthesis with the only difference of the solvent not being aqueous.
The merit of HT and ST methods is that it does not need very high synthetic temperature (normally below 500 K) [81,118], but the
synthesized products often have a high quality, which can show identical information of crystal structure. In particular, HT and ST
methods can realize precise control of shape, size, and crystallinity of nanostructures by altering the experimental parameters, such as
the reaction time, temperature, precursor type, solvent type, and surfactant type. In this respect, HT and ST methods are useful to
fabricate nanostructured products. In terms of the aqueous solution method [82,235], it is similar to HT/ST methods. The difference
is that for aqueous solution method, the boiling solution is always under normal pressure, and the reaction time is shorter than
traditional HT/ST methods.

4.2. Characterization

The characterization of single crystal SnSe is important to determine the compositions of SnSe, as well as verify the quality of their
crystal structures. Within these characterization methods, XRD is often used to determine the purity and structural features of the
obtained products, and SEM, TEM and scanning tunneling microscope (STM) [236] are frequently used to illustrate the morpholo-
gical, structural and compositional characteristics of the obtained products.

4.2.1. Macro-sized crystal


For macroscopic SnSe single crystals, to verify the crystal plane of synthetized SnSe crystals, XRD was used [61,216]. Fig. 20(a)
shows a typical α-SnSe cleavage plane of (4 0 0) [61], indicating that α-SnSe crystal is cleaved at the plane that is perpendicular to the
a-axis. Fig. 20(b) shows a typical EBSD image on a 1 mm2 surface of α-SnSe crystal [61], indicating homogeneous orientation within a
large area of sample surface with relatively homogeneous orientation. Fig. 20(c) shows a STM topographic image obtained from an in-
situ cleaved (0 0 1) surface of SnSe [97], in which a layered structure can be seen. Fig. 20(d) shows a HRTEM image of α-SnSe crystal.
The bottom inset of Fig. 20(d) is the corresponding selected area electron diffraction (SAED) pattern taken along the [0 1 1] zone-axis
while the top inset is the line profile along the dotted line AB in the main panel, which shows the d spacing of (1 0 0) atomic planes
[61]. All this information illustrates that such a single crystal is α-SnSe with a space group of Pnma at room temperature. Fig. 20(e)
and (f) show the SAED patterns obtained from the same TEM specimen area at room temperature and 820 K [61], respectively. As can
be seen, at room temperature, α-SnSe with a [2 1 1] zone-axis can be seen. When the temperature reached to 820 K, phase transfor-
mation took place and the resultant β-SnSe crystal showed a [1 2 1] zone-axis. Moreover, when the TEM specimen was cooled down to
room temperature, the SAED patterns transfer again from Fig. 20(f) to (e), indicating that the phase transition between α-SnSe and β-
SnSe is reversible.
In terms of the doped SnSe single crystals, Fig. 21(a) shows the XRD patterns of SnSe single crystals doped with different amount
of Na and Ag and their corresponding simulated diffraction patterns [61]. However, it is hard to see the peak deviation for the doped
single crystal, which is controversial. Fig. 21(b) shows a typical STM topographic image for SnSe on the b–c plane doped with Bi [52].
A few defects were found in the marked area by the black dashed circle and the inset HR-STM image. This distinctive feature is
predicted to be originated by substitutional Bi atom occupying the Sn-site [52]. Fig. 21(c) shows a TEM image of Na-doped SnSe

306
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 20. (a) XRD pattern of SnSe on the cleavage plane and the simulated diffraction pattern, (b) EBSD map, reproduced with permission [61],
Copyright 2014, Nature Publishing Group, (c) STM topographic image obtained from the in-situ cleaved (0 0 1) surface of SnSe, reproduced with
permission [97], Copyright 2016, Elsevier Publishing, (d) HRTEM image with bottom inset SAED pattern and top inset line profile along the dotted
line AB in the main panel showing the d spacing of (1 0 0), SAED patterns obtained at room temperature (e) and 820 K (f). Reproduced with
permission [61]. Copyright 2014, Nature Publishing Group.

crystals with stacking faults (red arrows) and dislocations (green arrows) can be observed [78]. Fig. 21(d) shows STEM image of the
typical regions of the Na-doped SnSe crystals [78]. The high magnification annular bright field (ABF) image is inserted in Fig. 21(d),
indicating the two stacking faults. Other studies showed that using ABF-STEM, interstitial atoms can be observed in Sn0.985Na0.015Se
single crystal [237], indicating that vast off-stoichiometric defects exist in single crystalline SnSe.

4.2.2. Micro/nano-sized crystal


Compared with macroscopic crystalline SnSe discussed above, to study the morphology of synthesized micro/nanoscopic crystals,
XRD, SEM and TEM were all commonly used. Fig. 22(a) shows the XRD patterns of SnSe crystals with different average size (∼30 and
∼100 μm, respectively) [91]. All diffraction peaks for both type of crystals can be exclusively indexed as the orthorhombic structured
α-SnSe with the lattice parameters of a = 11.37 Å, b = 4.19 Å, and c = 4.44 Å and space group of Pnma (Standard Identification Card,
JCPDS 48-1224) [91]. In Fig. 22(a), (4 0 0) peak is the strongest peak, suggesting that these single crystals contain significant {1 0 0}
surfaces. The inserted optical images of crystals show features of metallic luster. Fig. 22(b) shows the magnified XRD patterns of SnSe
crystals. As can be seen, both type of crystals have peak deviation to some extent, indicating that the self-doping exists in the crystals
[91]. The (1 1 1) peak for crystals with an average size of ∼30 μm was much stronger than that of others with a larger average size,
indicating that crystals with a larger average size possess much more {1 0 0} surfaces than their counterparts.
Fig. 22(c) is the typical SEM image of SnSe single crystals with an average of ∼30 μm and shows that the crystals have a plate-like
morphology [91]. The inserted image in Fig. 22(c) is a zoom-in SEM image taken from a typical plate, from which the thickness of the
plate can be estimated to be ∼3 µm [91]. Fig. 22(d) shows another SEM image of the plate-like SnSe single crystals with a larger
average size and a thickness of ∼5 µm [91]. Fig. 22(e) shows the plane information of one plate, indicating that the {1 0 0} planes are
the preferred surfaces [91]. Fig. 22(f) is a TEM image taken from a typical plate laid perpendicular to the electron beam [91].
Fig. 22(g) and 22(h) are the corresponding HRTEM image and the SAED pattern taken along the [1 0 0] zone-axis of the plate [91],
indicating typical orthorhombic structure of α-SnSe.

4.3. Performance

Because of the anharmonic bonding of Sn and Se, the thermoelectric properties possess totally different values along different axes
in single crystal SnSe [61]. Here, we summarize the main strength and drawback caused by anisotropic performance.

307
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 21. (a) XRD patterns of Sn1-xRxSe (R = Ag, Na; x = 0, 0.01, 0.02, 0.03) along the cleavage plane, reproduced with permission [79], Copyright
2015, Royal Society of Chemistry, (b) STM topographic image of Bi-doped SnSe with inserted high resolution image of a Bi dopant, reproduced with
permission [52], Copyright 2016, Nature Publishing Group, low-magnification image of the typical regions of the Na doped SnSe in (c) TEM model
and (d) STEM model with inserted inset high magnification ABF image, the red vertical arrows indicate contrast of stacking faults and the green
horizontal arrows indicate full/partial edge dislocations. Reproduced with permission [78]. Copyright 2015, American Association for the Ad-
vancement of Science.

4.3.1. Strong anisotropy


Along the b-axis, a maximum ZT of ∼2.6 was achieved at 923 K, derived from a high S2σ (∼8.5 μW cm−1 K−2) and a low κ
(∼0.35 W m−1 K−1) measured at the temperature [61]. It was found that an enhanced n (∼7.3 × 1018 cm−3 at 773 K) derived from
the thermal activation contributed to a high σ of ∼72.9 S cm−1 and a high S of ∼343.2 μV K−1 at 923 K, and the anharmonic
bonding of Sn and Se along the b-axis contributed to a low κ of ∼0.35 W m−1 K−1 [61]. κl (∼0.23 W m−1 K−1) occupied 65% of κ at
923 K, indicating that κ was dominated by lattice phonon transport along the b-axis [61]. Similarly, along the c-axis, a maximum ZT of
∼2.3 was achieved at 923 K, derived from a high S2σ of ∼6.6 μW cm−1 K−2 and a low κ of ∼0.30 W m−1 K−1 measured at the
temperature, both of which were lower than that measured along the b-axis [61]. However, along the a-axis, a maximum ZT of only
∼0.8 was obtained at 923 K, derived from a low S2σ (∼1.9 μW cm−1 K−2) and a low κ (∼0.23 W m−1 K−1) measured at the
temperature, both of which were much lower than that measured along the b- and c-axes [61]. This is because along the a-axis, the
SnSe layers arrayed along the a-axis composed by weak van der Waals force resulted in a much lower μ (∼7.2 cm2 V−1 s−1 at 773 K),
and in turn resulted in a much lower σ (∼12.8 S cm−1 at 923 K) [61].

4.3.2. Strength and drawback


Although there are strong anisotropy for the thermoelectric properties of single crystal SnSe, the properties still possess the same
trend and show extraordinary performance at high temperature. Fig. 23 shows the main thermoelectric properties (σ, S, S2σ, n, μ, κ, κl,
κl/κ and ZT) of SnSe single crystals along the a-, b- and c-axes discussed above [61]. For σ, as shown in Fig. 23(a), from 300 to 525 K,
an obvious decrease showed a metallic transport behavior of σ [61]. From 525 to 800 K, the σ value increased by the function of
carriers thermal excitation, which started to show a thermally activated semiconducting behavior of σ [61,238]. After that, the σ
value slightly decreased again due to the phase transition from α-SnSe to β-SnSe. For S, as shown in Fig. 23(b), the peak S values are
found between 500 K and 600 K due to the bipolar transport occurs [181]. Fig. 23(c) shows the maximum value of S2σ along the b-, c-
and a-axes were 10.1 µW cm−1 K−2, 7.7 µW cm−1 K−2 and 2.1 µW cm−1 K−2 at 850 K, respectively. Fig. 23(f) demonstrates that κ

308
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

30.0 30.5 31.0 31.5 32.0

a 0.99

0.66
size: ~ 30 μm size: ~ 100 μm
b 0.99

0.66

(400)
0.33 0.33
0.00
Intensity (a.u.) 0.00

Intensity (a.u.)

(400)
0.99 0.99

(111)

(800)

(111)
0.66 0.66
0.33 0.33
~ 100 μm
0.00 0.00

(400)
0.99 0.99

(400)
(800)

(111)
0.66 (111) 0.66
0.33 0.33
0.00
~ 30 μm
0.00
(111)

0.99 0.99
PDF#48-1224 PDF#48-1224

(111)
(400)
0.66 0.66

(400)
0.33 0.33
Peak deviation
0.00 0.00

20 25 30 35 40 45 50 55 60 65 70 30.0 30.5 31.0 31.5 32.0


2 theta (degree) 2 theta (degree)

Fig. 22. (a) XRD patterns of SnSe single crystals (blue line for crystals with an average size of ∼30 μm and orange line for crystals with an average
size of ∼100 μm) with inserted optical images, (b) magnified XRD patterns of SnSe single crystals to see the peak deviation, SEM image of SnSe
single crystals with inserted SEM image shows a typical plate for (c) crystals with an average size of ∼30 μm and (d) crystals with an average size of
∼100 μm, (e) description of planes for one SnSe single crystal plate, (f) TEM image of a typical SnSe single crystal plate, (g) corresponding HRTEM
image and (h) corresponding SAED pattern taken from the plate shown in (f). Reproduced with permission [91]. Copyright 2018, Elsevier Pub-
lishing.

decreased gradually before 800 K, then kept at a low value until 973 K, caused by the phase transition from α-SnSe to β-SnSe [61]. At
room temperature, the κ values were measured as 0.46 W m−1 K−1, 0.70 W m−1 K−1 and 0.68 W m−1 K−1 along the a-, b- and c-axes,
respectively, which are much lower than that of the other thermoelectric materials [3,5,239]. To calculate κl, κl = κ − κe = κ − LσT,
where L is the Lorenz number [240], which was approximated to 1.5 × 10−8 W Ω K−2 in the case of non-degenerate semiconductors
limited by phonon scattering, such as SnSe [241]. In fact, the theoretically calculated minimal κ are slightly larger than those from
experimental measurements, which could possibly result from the value of the L (1.5 × 10−8 W Ω K−2), which can vary in the range
of 1–2.4 × 10−8 W Ω K−2 [61]. The ratio of κl to κ shown in Fig. 23(h) is up to 65%, indicating that that κ was dominated by lattice
phonon transport (κl) [61].
Fig. 23(i) illustrates ZT values of single crystal SnSe along the a-, b- and c-axes. Although the ZT values are very high at high
temperature, however, they were low before 800 K, indicating that the ZT value obtained from α-SnSe was much lower than that of β-
SnSe. Considering that the mid-temperature (from 300 K to 700 K) is the most commonly used in practical applications, and the
mechanical property of β-SnSe is much weaker than that of α-SnSe, α-SnSe should possess much better potential for practical
application than β-SnSe. Thus, it is necessary to further improve the thermoelectric performance of α-SnSe.

309
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 90 a-axis b 600 c 12
a-axis
b-axis b-axis

S2σ (μW cm-1K-2)


550 9
σ (S cm-1) c-axis c-axis
60

S (μV K-1)
500 α-SnSe β-SnSe
6
α-SnSe β-SnSe 450 α-SnSe β-SnSe
30
400 a-axis 3
b-axis
350
0 c-axis 0
300
300 450 600 750 900 300 450 600 750 900 300 450 600 750 900
T (K) T (K) T (K)

d a-axis e 250 a-axis f 1.0 a-axis


b-axis b-axis b-axis

κ (W m-1K-1)
1019 200 c-axis
c-axis μ (cm2 V-1s-1) c-axis 0.8
150 β-SnSe
n (cm-3)

α-SnSe
0.6
100
1018
0.4
50

0 0.2
1017 300 400 500 600 700 800 300 450 600 750 900
300 400 500 600 700 800
T (K) T (K) T (K)
g 1.0 a-axis h 1.0 i 2.5 a-axis
b-axis b-axis
2.0
κl (W m-1K-1)

0.8 c-axis c-axis


0.9
1.5
κ /κ

α-SnSe β-SnSe
ZT
0.6 α-SnSe β-SnSe
0.8
l

1.0
0.4 a-axis
0.7 0.5
b-axis
0.2 c-axis 0.0
0.6
300 450 600 750 900 300 400 500 600 700 800 300 450 600 750 900
T (K) T (K) T (K)
Fig. 23. Temperature-dependent thermoelectric properties for single crystal SnSe: (a) σ, (b) S, (c) S2σ, (d) n, (e) μ, (f) κ, (g) κl, (h) κl/κ and (i) ZT along
different axes. The phase transformation is pointed out by orange dash line. Reproduced with permission [61]. Copyright 2014, Nature Publishing
Group.

4.4. Improvement

In order to further enhance the thermoelectric properties of α-SnSe, doping was used to realize p-type or n-type single crystals
[78,79]. For p-type doping, the dopants used include alkali metals (such as Na) [78,79], and I-B group metals (such as Ag) [79,242].
For n-type doping, the dopant used was V-A group metals (such as Bi [52]). Most of the dopants were used to substitute Sn positions
to tune n.

4.4.1. p-type doping


For p-type doping, a maximum ZT of ∼2.0 was achieved via doping with 3% Na at 800 K along the b-axis [79]. Such significant
thermoelectric performance derived from the obviously enhanced n values [79]. Experiments demonstrated that, with increasing the
Na content, n increased from ∼3.2 × 1018 cm−3 [61] to ∼2.1 × 1019 cm−3 doped with 1% Na [79] and ∼2.7 × 1019 cm−3 doped
with 2% Na at 723 K [79], respectively. Although a high n resulted in a lower S value (∼261.4 μV K−1) than that of undoped crystals
(∼437.0 μV K−1) at 773 K [61], a much higher σ (∼212.1 S cm−1) [79]contributed to a much larger S2σ (∼14.5 μW cm−1 K−2) than
that of single crystals at the same temperature [79]. After doping with 3% Na, κ increased from 0.33 W m−1 K−1 [61] to 0.58 W m−1
K−1 [79] at 773 K. Other experiments also achieved a maximum ZT of ∼2.0 at 773 K along the b-axis in hole-doped SnSe via 1.5% Na
[78]. The significant thermoelectric performance also raised from high S2σ (∼14.2 μW cm−1 K−2), which came from high σ (∼149.5
S cm−1) and strongly enhanced S (∼307.8 μV K−1) at 773 K [78]. By angle-resolved photoemission spectroscopy, large hole effective
masses were revealed in both undoped and hole-doped SnSe crystals [243], contributing to the high electrical transport performance.
For Ag-doped SnSe single crystals [79,242], a maximum ZT of ∼1.0 was obtained at 800 K along the b-axis via doping with 3% Ag
[79]. Compared with doping with 3% Na [79], a lower σ (∼71.6 S cm−1) resulted in a lower S2σ (∼7.86 μW cm−1 K−2) for doping

310
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 1500
b 600
Pure
3 % Na
1200 3 % Ag 300

σ (S cm-1)

S (μV K-1)
1.5 % Na
900
6 % Bi 0 Pure 3 % Na
600 3 % Ag 3 % Ag 1.5 % Na
6 % Bi 3 % Ag
-300
300

0 -600
300 400 500 600 700 800 300 400 500 600 700 800
T (K) T (K)
c 50 Pure 3 % Na d 2.5 Pure
3 % Ag 1.5 % Na 3 % Na
S2σ (μW cm-1K-2)

40 6 % Bi 3 % Ag 2.0 3 % Ag

κ (W m-1K-1)
1.5 % Na
30 1.5 6 % Bi
3 % Ag
20 1.0

10 0.5

0 0.0
300 400 500 600 700 800 300 400 500 600 700 800
T (K) T (K)
e 3.0 Pure 3 % Na f 3.0 Average
3 % Ag 1.5 % Na Peak
2.5 6 % Bi 3 % Ag 2.5
2.07 2.06
2.0 2.0 1.94
ZT

ZT

1.52 1.55
1.5 1.5
1.12
0.96 1.02
1.0 1.0
0.55
0.5 0.5 0.43 0.44
0.3

0.0 0.0 Pure 3 % 3 % 1.5 % 6 % 3 %


300 400 500 600 700 800
Na Ag Na Bi Ag
T (K)
Fig. 24. Temperature-dependent thermoelectric properties for pure SnSe [61], Sn0.97Na0.03Se [79], Sn0.97Ag0.03Se [79], Sn0.985Na0.015Se [78], 6%
Bi-doped SnSe [52], and Sn0.97Ag0.03Se [242]: (a) σ, (b) S, (c) S2σ, (d) κ and (e) ZT. (f) Comparison of both average and peak ZT between different
materials from 323 to 773 K.

with 3% Ag, and a higher κ (∼0.64 W m−1 K−1) was obtained at the same time [79]. Compared with Ag, the significant mass and size
difference of Na and Sn atoms gave rise to a strong phonons scattering, and in turn contributed to a low κ [79].

4.4.2. n-type doping


For n-type doping, a maximum ZT of ∼2.2 was achieved at 733 K along the b-axis via 6% Bi-doping with n of −2.7 × 1019 cm−3
[52,244]. The excellent thermoelectric properties of n-type SnSe single crystals can be attributed to that σ (∼61.5 S cm−1) is
enhanced via the Bi doping while maintaining a low κ (∼0.29 W m−1 K−1) [52]. It demonstrated that both n- and p-types in a one-
material system can be realized, and with both excellent performance.
Fig. 24 plots the main thermoelectric properties (σ, S, S2σ, κ and ZT) of SnSe single crystals before and after doping with different
dopants along the b-axis [52,78,79] (except doping with 3% Ag [242] with pink line, which is along the a-axis). Theoretical cal-
culations showed that a peak ZT of 2.57 can be obtained in p-type SnSe along the b-axis by optimizing n to ∼3.6 × 1019 cm−3 [196].
Theoretical calculations indicated that heavy-light band overlapping near the valence band and a mixed ionic-covalent bonding were
the reasons for this extraordinary thermoelectric performance [196]. For n-type doping, theoretical calculations predicted that a peak
ZT of ∼3.1 can be obtained for n-type SnSe along the a-axis at 770 K at n of 2.8 × 1019 cm−3 [204]. Fig. 24(f) shows the comparison
of both average and peak ZT between doped and undoped single crystal SnSe from 300 to 773 K. As can be seen, both Na-doped p-
type single crystals and Bi-doped n-type single crystals have much higher average and peak ZT than those of their undoped

311
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Table 5
Comparison of calculated/measured ρ and κ for SnSe crystals along different directions reported in the literature (ρth = 6.18 g cm−3). The measured
ρ is labelled with *.
Product axis κ at ∼323 K D at ∼323 K Cp at ∼323 K ρ at ∼323 K κ at ∼773 K D at ∼773 K Cp at ∼773 K ρ at ∼773 K Ref.
(W m−1 K−1) (mm2 s−1) (J g−1 K−1) (g cm−3) (W m−1 K−1) (mm2 s−1) (J g−1 K−1) (g cm−3)

Pure a ∼0.45 ∼0.33 ∼0.252 ∼5.45 ∼0.23 ∼0.16 ∼0.259 ∼5.51 [61]
Pure b ∼0.68 ∼0.49 ∼0.252 ∼5.43 ∼0.33 ∼0.23 ∼0.259 ∼5.49 [61]
Pure c ∼0.65 ∼0.47 ∼0.252 ∼5.44 ∼0.28 ∼0.20 ∼0.259 ∼5.40 [61]
3% Na b ∼1.70 ∼1.08 ∼0.252 ∼6.25 ∼0.58 ∼0.34 ∼0.259 ∼6.23 [79]
3% Ag b ∼1.92 ∼1.21 ∼0.252 ∼6.28 ∼0.64 ∼0.40 ∼0.259 ∼6.20 [79]
1.5% Na a ∼0.62 ∼0.40 ∼0.250 ∼6.15, ∼0.27 ∼0.16 ∼0.265 ∼6.43 [78]
∼6.0*
1.5% Na b ∼1.55 ∼1.00 ∼0.250 ∼6.19, ∼0.54 ∼0.35 ∼0.265 ∼5.89 [78]
∼6.0*
1.5% Na c ∼1.29 ∼0.82 ∼0.250 ∼6.25, ∼0.40 ∼0.25 ∼0.265 ∼5.99 [78]
∼6.0*
6% Bi a ∼0.46 ∼0.30 ∼0.252 ∼6.10 ∼0.18 ∼0.12 ∼0.259 ∼6.01 [52]
6% Bi b ∼0.82 ∼0.53 ∼0.252 ∼6.13 ∼0.27 ∼0.17 ∼0.259 ∼6.19 [52]
6% Bi c ∼0.73 ∼0.47 ∼0.252 ∼6.12 ∼0.24 ∼0.15 ∼0.259 ∼5.99 [52]
3% Ag a ∼2.05 – ∼0.252 6.178* ∼0.61 – ∼0.259 [242]

counterparts in this temperature range. These results indicate that doping is a promising method to enhance the thermoelectric
performance of α-SnSe, as well as a convenient way to realize n-type thermoelectric materials with a high thermoelectric perfor-
mance.

4.5. Controversy on thermal conductivity

Table 5 summarizes the measured density (ρ) by using Archimedes’ method and estimated ρ of SnSe single crystal along different
axes if the original articles have not provided, in which ρ was estimated using the relation [61]:
κ = DCp ρ (10)
where D is the thermal diffusivity, and Cp is the specific heat capacity at constant pressure. Experimentally, measured D, Cp, and κ can
be used to determine ρ. For the reported undoped SnSe single crystal [61], the calculated ρ was only ∼5.45 g cm−3 along the three
axes, which was much lower than the theoretical value (6.18 g cm−3), indicating a very low relative density (∼88%). In this respect,
the observed ultralow κ values may not be intrinsic for fully dense single-crystalline SnSe (chemical ratio is SnSe = 1:1) [245].
However, considered that the SnSe single crystal has a strong anisotropy and weak mechanical properties which lead to facile
cleavage, the crystal can easily be damaged during cutting, handling and measurements, and in turn result in a high uncertainty
between 15% and 20% for the measured κ [61]. In this respect, the calculated low ρ is reasonable. At the same time, experiments
based on a vertical-Bridgman-grown single crystal SnSe [246] with a high density (ρ = 6.10 ± 0.05 g cm−3) showed that the intrinsic
κ of single crystal SnSe is likely significantly higher than that of the reported undoped SnSe single crystals with a density of ∼88%
[61]. The single crystal SnSe with a relative density of 99% showed that the κ values along the a-, b-, and c-axes were 1.2, 2.3, and
1.7 W m−1 K−1 at 300 K, respectively [246], which were higher than the values of 0.46, 0.70 and 0.67 along the a-, b-, and c-axes
[61]. These results were consistent with early studies performed on single crystal SnSe (1.9 W m−1 K−1 at 300 K perpendicular to the
c-axis) [247]. Besides, the single crystal SnSe with a relative density of 99% demonstrated that the κ values along the a-, b-, and c-axes
were all ∼1.0 W m−1K−1 at 700 K [246], which were higher than the reported results of ∼0.24, 0.34 and 0.31 along the a-, b-, and c-
axes, respectively [61]. In this respect, the reported single crystal with a low calculated ρ should be at least “SnSe crystal”.
For doped single crystal SnSe, the measured/calculated ρ values are also listed in Table 5. Compared with the case of undoped
single crystal, the 3% Na-doped single crystal possessed a calculated ρ value of ∼6.25 g cm−3, close to the theoretical value
(6.18 g cm−3) [79], which can be treated as within the uncertainty of measured results [79]. Similarly, the 1.5% Na-doped single
crystal possessed calculated ρ values of ∼6.43, ∼5.89 and ∼5.99 g cm−3 along the a-, b-, and c-axes at ∼773 K, respectively [78],
which derived from the κ uncertainty (∼12%) comprised uncertainties of 5% for the measured D, 5% for the measured Cp, and 2% for
the measured ρ (∼6.0 g cm−3). Although there exist controversies on the intrinsic κ values of single crystal SnSe, it should be noted
that SnSe-based thermoelectric material is still a promising candidate with full potentials for further improving the thermoelectric
performance.

5. Polycrystalline SnSe

Although SnSe single crystals have demonstrated excellent thermoelectric performance due to their ultralow κ along the b- and c-
axes [61], they suffered from the poor mechanical properties, rigid crystal growth conditions, and high cost of their production. Also,
the techniques used for growing single crystals are limited for industrial scale-up [65]. For this reason, SnSe single crystals may be
difficult to be employed in large-scale practical thermoelectric devices. To overcome this challenge, polycrystalline SnSe has become
a research topic, and has been preliminarily applied into the thermoelectric modules [248].

312
Table 6
A comprehensive summary on the synthesis of polycrystalline SnSe. Here, for the synthetic methods of SnSe, melting method is abbreviated as M, arc-melting method is abbreviated as AM, annealing is
abbreviated as A, mechanical alloying is abbreviated as MA, solid-state reaction is abbreviated as SSR, combustion method is abbreviated as CM, and aqueous solution is abbreviated as AS; For the
sintering methods, spark plasma sintering is abbreviated as SPS, hot-pressing is abbreviated as HP, cold-pressing is abbreviated as CP, and open die pressing is abbreviated as ODP.
Z.-G. Chen et al.

Product Type Dopant or Synthetic Synthetic pressure Synthetic Synthetic time Sintering Sinter Sinter Sinter time Year Ref.
mixture method (Torr) temperature (K) (h) method pressure temperature (K) (min)
(MPa)

SnSe p – M – 1223 24 SPS 50 698, 708 7 2014 [65]


SnSe p – AM – – – – – – – 2015 [88]
SnSe p – M 10−5 1173 18 HP 60 773 20 2016 [66]
SnSe p – M 10−4 1193 2 HP, SPS 55 753 5 2016 [89]
SnSe p – SSR 10−5 773 96 SPS 60 773 2 2016 [90]
SnSe p – M – 1200 12 SPS 60 830–850 – 2015 [249]
SnSe p – HT – 453 12 SPS 45 673 15 2015 [118]
SnSe p – ST – 393, 413, 433 12 SPS 50 773 5 2016 [81]
SnSe p – AS – – 2 HP 60 773 20 2016 [82]
SnSe p – M ∼7.5 × 10−6 1223 24 HP 600 723 60 2017 [252]
SnSe p – – – – – ODP – – – 2017 [251]
SnSe p – AM – – – – – – – 2017 [422]
Sn0.98Se p – ST – 503 36 SPS 60 850 5 2018 [91]
SnSe – – ST, A – – – – – – – 2005 [83]
SnSe p – ST – 473 24 HP 50 923 10 min 2017 [233]
SnSe p – HT – 423–443 6–12 CP + HP 2 tons 853 – 2017 [423]
SnSe p – – – – – CP, A – – – 2016 [250]
SnSe p – MA – – 20 SPS 30, 60, 80, 823 5–10 2017 [424]

313
100, 120
SnSe P – CM – – <1 SPS 50 773 5 2018 [150]
Sn1-xNaxSe p Na M < 10−6 1223 12 SPS – 746 20 2016 [161]
Sn1-xNaxSe p Na M 7.5 × 10−5 1253 24 SPS 50 873 5 2016 [274]
Sn1-xNaxSe p Na M ∼3 × 10−4 1173–1223 6 HP 50 873 7 2016 [258]
Sn1-xNaxSe p Na – – – – SPS 40 873 5 2017 [425]
Sn1-xNaxSe p Na High pressure – 1123 0.5 + 1 SPS 70 823 10 2017 [285]
Sn1-xNaxSeClx p NaCl SSR – 1073 10 min Rapid HP 50 ∼823 5 2017 [426]
NaySn1−xPbxSe p Na, Pb M ∼10−4 1223 12 SPS 50 ∼783 5 2017 [286]
SnSe1-xTex p Te SSR – 1023 1 HP 80 – 120 2012 [267]
SnSe1-xTex p Te MA (450 rpm) – – 15 SPS 50 793 5 2015 [128]
SnSe1-xTex p Te Microwave-ST – 503 6 SPS 40 873 5 2017 [232]
AgxSn1-xSe p Ag M, A 7.5 × 10−5 1253 24 SPS 50 873 5 2016 [277]
AgxSn1-xSe p Ag M, A – 1200 12 HP 36 800 10 2014 [268]
AgxSn1-xSe p Ag ST – 413 15 SPS 50 500 5 2017 [288]
Sn1-xTlxSe p Tl M ∼7.5 × 10−6 1223 6 HP 70 673 60 2016 [278]
SnSe p Al, Pb, In, Cu A – 773 ∼72 SPS 60 773 ∼10 2016 [269]
Sn1-xCuxSe p Cu M – 1203 2 HP 6000 1073 3 2017 [287]
Sn1-xCuxSe p Cu HT – 403 36 SPS 50 693 7 2018 [296]
SnSe1-xInx p In M ≤10−5 1223 48 HP 70 850 60 2016 [279]
Sn1-xGexSe p Ge AM – – – – – – – 2016 [289]
Sn1-xGexSe p Ge ZM ∼7.5 × 10−5 1223 2 HP 60 753 20 2017 [290]
Sn1-xGexSe p Ge M – 1203 2 – – – – 2017 [291]
KxSn1-xSe p K MA (425 rpm) – – 5, 8, 12 SPS 50 773 5 2016 [67]
SnSe p Li, Na, K M 10−4 1223 8 SPS – 873 5 2016 [283]
NaxKySn1-x-ySe p Na, K MA (425 rpm) – – 5 SPS 50 773 5 2017 [270]
Progress in Materials Science 97 (2018) 283–346

(continued on next page)


Table 6 (continued)

Product Type Dopant or Synthetic Synthetic pressure Synthetic Synthetic time Sintering Sinter Sinter Sinter time Year Ref.
mixture method (Torr) temperature (K) (h) method pressure temperature (K) (min)
Z.-G. Chen et al.

(MPa)

ZnxSn1-xSe p Zn M ∼7.5 × 10−6 1223 24 HP 600 673 60 2016 [262]


Sn1-xSe p SnSe2 M, A 10−3 1223 24 SPS 50 708 6 2017 [427]
Ag0.015Sn0.985Se p Ag M – 1273 10, 8 – – – – 2017 [275]
Sn1-xSmxSe p Sm M – 1203 2 Rapid HP 6000 1073 3 2017 [292]
SnSe p LaCl3 MA (425 rpm) – – 8 SPS 50 773 5 2018 [293]
SnSe0.985Cl0.015 p SnCl2 M – 1273 10, 8 – – – – 2017 [275]
(Cu2Se)1-x(SnSe)x p β-Cu2Se M – 1403 10 SPS 50 923 5 2015 [303]
SnS1-xSex p SnS M, A 10−4 1223 20 SPS 50 903 7 2015 [129]
SnSe-SnS p SnS MA (450 rpm) – – 15 SPS 50 903 7 2017 [299]
Sn1-xPbxSe p PbSe HT – 403 36 SPS 50 693 7 2016 [84]
(Sn1-xPbx)0.99Na0.01Se p Na, PbSe M, A – 1223 10 SPS 50 873 5 2017 [284]
SnTe/Ag0.005Sn0.995Se p SnTe, Ag M ∼7.5 × 10−6 1223 24 SPS 20 673 20 2016 [276]
composites
SnSe/C composites p Carbon black M – 1223 10, 24 HP 600 673 60 2016 [298]
SnSe/PbTe composites p PbTe M ∼7.5 × 10−6 1223 10, 24 HP 600 673 ∼96 2016 [301]
SiC/SnSe composites p SiC ceramics M – 1223 10 HP 50 823 10 2016 [260]
SnSe/MoS2/ graphene p MoS2/graphene SSR, HT – ∼1073 1/6 HP 50 ∼823 5 2017 [85]
composites
SnSe/MoSe2 composites p MoSe2 M ∼7.5 × 10−7 1223 24 SPS 50 773 5 2017 [300]
Ag0.01Sn0.99Se1−xSx p Ag, SnS M, A – 1273 24 HP 50 773 30 2017 [302]
Na0.015Sn0.985Se/CNTs p Na, carbon M, A – 1253 24 SPS 50 753 5 2018 [305]

314
composites nanotubes
SnSe p Se nanowires MA (150 rpm), – 453 1 HP 50 823 10 2018 [297]
HT
SnS0.2Se0.8 p Ag MA (450 rpm) – – 12 SPS 50 813, 833 and 853 6 2018 [306]
SnSe1−x n – ST – 453 24 – – – – 2016 [86]
SnSe1-x n – M ∼3.75 × 10−6 1273 20 HP 60 923 60 2018 [282]
SnSe n Cl AS – – 2 HP ∼60 773 20 2017 [235]
SnSe0.95-BiCl3 n BiCl3 M ∼7.5 × 10−6 1193 1 SPS 55 753 8 2016 [281]
SnSe0.95-PbBr2 n PbBr2 M ∼7.5 × 10−5 1193 1 HP 60 753 30 2017 [294]
SnSe1-xSx n SnS, I M – 1193 6 HP 50 873 7 2016 [259]
Pb1-xSnxSe n PbSe MA (450 rpm) – – 24 SPS 50 873 5 2015 [257]
Sn1-xTixSe, Sn0.94-xPbxTi0.06Se n Ti, Pb MA (450 rpm) – – 6 SPS 50 773 5 2017 [295]
Sn1-xXxSe – Ca, Sr, Ba, Cu, M, A 10−3 1273 18 HP 100 798 45 2015 [127]
Zn, Ge
Progress in Materials Science 97 (2018) 283–346
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

5.1. Fabrication

Texturing and doping have been the two key approaches used for improving thermoelectric performance of polycrystalline SnSe,
from which their ZT values have been improved from ∼0.4 to ∼1.7, although such ZT values are still lower than their single crystal
counterparts, mainly due to their relatively high κ and low σ [65,66,81,88,89,118,119,249–252]. Table 6 summarizes the synthesis
methods with relevant parameters for producing polycrystalline SnSe materials.

5.1.1. Synthesis of product


As shown in Table 6, the melting [225–227] or solid-state reaction [228–230] routes are the two most commonly used methods,
especially for the melting route. The principles and main procedures of the melting and solid-state reaction routes have been dis-
cussed in Section 4.1. For the arc-melting route [88,253], it is different from the traditional melting route, because the heats for arc-
melting are provided through an electric arc [254]. The precursors are directly exposed to the electric arc, and the current in the
furnace terminals passes through the charged precursors. In this regard, the arc-melting has a much higher heating speed than the
traditional melting.
When the melting or solid-state reaction were finished, sometimes the annealing, keeping for several days and then cooling down
to room temperature [249,250], followed at the temperature below the melting point of SnSe to change the ductility and hardness of
the obtained SnSe [255]. Moreover, the annealing can lead to re-alignment of crystalline domains and in turn result in the increase in
S and σ [250].
Mechanical alloying [256] is a solid-state powder processing technique. By blending SnSe powder precursors in a high-energy ball
mill, it is powerful to produce a homogeneous SnSe product. During mechanical alloying, the precursors were milled on a planetary
ball mill at an appropriate speed for an appropriate time [257] by using a stainless steel vessel filled with a mixed atmosphere such as
Ar and H2 gases. Sometimes, mechanical alloying was also treated as a step after melting [258–260]. Experiment showed anisotropy
in the grains can be achieved by ball milling for 1 min [258].

5.1.2. Sintering
In order to obtain bulk SnSe materials, the synthesized SnSe products should be ground into powders and then sintered, to realize
a polycrystalline structure with high density and a good mechanical property. Because the polycrystalline SnSe fabricated via cold-
pressing route has demonstrated a poor thermoelectric performance [250], while hot-pressing and SPS have been demonstrated as
the two most commonly used sintering methods to fabricate polycrystalline SnSe bulk materials. Traditional hot-pressing is a low-
strain-rate powder metallurgy process taken under high pressure, commonly used to produce brittle and hard materials [261]. For
SnSe, hot-pressing can sinter SnSe powder compactly at a high temperature and a high pressure (up to 600 MPa) [252,262].
In terms of the SPS technique, it has played a dominant role in the sintering of powder compacts with high density at lower
sintering temperature compared to conventional sintering techniques such as hot or cold pressing [263]. It is different from the
traditional hot-pressing for that the heat generation is internal during SPS process, and the heat is provided by external heating
elements during hot-pressing. For SnSe, SPS can achieve a very high heating and cooling rate (up to 1000 K min−1), resulting in a
very fast sintering process (within several minutes) [150]. SPS has the great potential of densifying SnSe powders with high density
and orientation. After sintering, annealing treatment can be used to re-alignment the crystal crystalline domains of SnSe pellets
[250].

5.2. Feature

The characterization of polycrystalline SnSe is important to confirm the composition of SnSe for both synthesized products (such
as powders) and sintered bulks. Within these characterization methods, XRD and XPS are used to analyze the purity and binding
structures of SnSe products, and XRD is also used to verify the anisotropy of sintered bulks. SEM and TEM are frequently used to
characterize the morphology and structural information of polycrystalline SnSe.

5.2.1. Phase analysis


Fig. 25(a) shows a typical XRD pattern obtained from SnSe synthesized via solid-state reaction [90]. All the peaks can be indexed
to the orthorhombic phase of α-SnSe with a strong (4 0 0) diffraction peak. The average crystallite size of the as-synthesized SnSe was
determined to be ∼42 nm using Williamson–Hall method [90]. Apart from XRD, XPS can also be used to analyze the phase of SnSe
[81,264–266]. Fig. 25(b) shows a partial high resolution XPS spectrum of SnSe synthesized via solvothermal route [81], indicating
the existence of one valence state for Sn and Se. The peaks of Sn 3d3/2 and Sn 3d5/2 were singlet and no accessorial binding energy
peaks can be found, indicating the divalent characteristic of Sn ions, and a binding energy peak of Se 3d was also observed at 53.7 eV,
suggesting Se2+ [81].
To determine the compositional ratio of Sn and Se in the materials, EDS [267–269] and high-resolution electron probe micro-
analysis (EPMA) [81,91,128,257] are commonly used. Fig. 25(c) shows a typical EDS spot analysis of SnSe synthesized by micro-
wave-assisted solvothermal route, the atomic ratio of Sn and Se was roughly 1:1. To exam the existence and dispersion of dopants,
EDS mapping is also frequently used. Fig. 25(d) shows an EDS mapping of Ag in Ag-alloyed SnSe and the observed contrast indicates
the successful alloying of Ag [268] and that Ag is segregated in some regions of few microns. To obtain more exact stoichiometry of
the as-synthesized SnSe products, EPMA is widely used. A typical result is shown in Sn0.98Se, which provide the stoichiometric ratios
of Sn:Se = 49.5:50.5 ± 0.4 for self-doped SnSe [91].

315
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a (400) SnSe b Se
3d5/2 Sn
Se 3d

Intensity (a.u.)
Intensity (a.u.)
3d3/2

(111) Sn 3d
(110)
(212) (800)
(201) (020)

10 20 30 40 50 60 70 80 50 55 60 490 500
2 theta (degree) Binding Energy (eV)
c 500
Se At%
400 Sn 50.1
Sn
Intensity (a.u.)

300 Se 49.9

200

100
Sn Sn Se
Se
0
0 5 10 15 20
keV
Fig. 25. (a) XRD pattern of SnSe powders via solid-state reaction [90], (b) XPS spectrums of SnSe synthesized by solve-thermal route [266], (c) EDS
spot spectrum of pure SnSe synthesized by microwave-assisted solve-thermal route, and (d) EDS mapping image of 1% Ag alloyed SnSe alloys.
Reproduced with permission [268]. Copyright 2014, Royal Society of Chemistry.

5.2.2. Sinter-induced anisotropy


Because SnSe has a typical layered structure, anisotropy in the sintered polycrystalline SnSe materials is widely existed [66]. To
verify the anisotropy, XRD and SEM are used. Fig. 26(a) is the typical XRD patterns at the direction perpendicular to the sintering
pressure (marked as ⊥) and parallel to the sintering pressure (marked as //) for the sintered polycrystalline SnSe [91], respectively.
As can be seen, the pellet cut along the ⊥ direction shows a strong (4 0 0) peak, similar to the XRD pattern taken from plates (refer to
Fig. 22(a)) [91]. In contrast, the pellet cut along the // direction shows a weak (4 0 0) peak, but a strong (1 1 1) peak, indicating a
strong anisotropy in the sintered SnSe pellets. Fig. 26(b) is the typical SEM images of the polished pellet surface cut along the ⊥
direction [91], indicating that the flat surface is without observable flaw. Fig. 26(c) and (d) are the typical SEM images of pellets
fractured along ⊥ and // directions to the SPS pressure [91] respectively, from which distinct fracture features confirm that the
sintered pellets have a preferred orientation. Fig. 26(e) is a HRTEM image taken from the pellet and the inserted SAED pattern taken
along the [0 0 1] zone-axis of the pellet, indicating the typical orthorhombic structure of SnSe.

5.2.3. Microscopic study


Fig. 27(a)–(c) shows the high-resolution high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM)
images of polycrystalline SnSe viewed along the a-, b- [270], and c-axis [270] with overlays showing Sn atoms in red and Se atoms in
blue, respectively. The dashed rectangle indicates the projected unit cell. Fig. 27(d) shows an atomic-scale EDS mapping from zone
axis of [0 1 0] [270], confirming the arrangement of Sn and Se atoms. For the case of sintered bulks with nanoprecipitates, Fig. 27(e)
shows an EBSD phase map of SnSe with grains identified as cubic rock-salt phase in red and orthorhombic α-phases in blue [141], and
Fig. 27(f) shows a TEM image of SnSe with rock-salt phase with an inserted SAED pattern detected on the grain [141], indicating the
existence of rock-salt phase of SnSe nanoprecipitates as second phase. Experimental and theoretical studies have shown that cubic
rock-salt SnSe has an almost negligible energy difference from α-phase [271,272]. Possible transition from α-SnSe to cubic SnSe
phase will occur due to a strong correlation between the lattice constants of isostructural and chemically analogous cubic phase and
α-phase. Recent studies showed that the cubic rock-salt SnSe can be induced by microstructure control [273].
TEM is a powerful tool to determine the nature of crystal defects. Fig. 27(g) is a HRTEM image taken from a sintered SnSe pellet
[119], and shows nanosized grains with disordered atomic planes with dislocations (marked by dashed circle) and lattice curvatures
(marked by dashed line) [119]. For doping with other elements (such as Na [274] and Ag [268]), Fig. 27(h) shows a HRTEM image of
Ag-doped SnSe [275], from which a dark area was observed to be predicted to be amorphous. In terms of incorporating with other
compounds (such as SnTe), Fig. 27(i) shows a TEM image of the boundary of two phases for SnSe and SnTe [276]. Besides, TEM is
also useful to characterized nanoprecipitates [67,141], as discussed in Section 2.4.

316
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 0.99
P

(400)
0.66
X
X-Ray

(111)
0.33 S urface pressure
0.00
Intensity (a.u.)

(111)
0.99

X-Ray
X

(400)
0.66

0.33 Surface pressure


0.00
(111)

0.99 (400)
0.66
PDF#48-122
24
0.33

0.00

20 25 30 35 40 45 50 55 60 65 70
2 theta (degree)

Fig. 26. (a) XRD patterns of sintered SnSe pellets measured along both the ⊥ (red line) and // (green line) direction, (b) SEM image of the surface of
pellets polished along the ⊥ directions, SEM images of pellets fractured from (c) the ⊥ directions and (d) the // directions, (e) HRTEM image of
pellets. Reproduced with permission [91]. Copyright 2018, Elsevier Publishing.

5.3. Texturing

As mentioned in Section 5.1, polycrystalline SnSe has become a research topic due to their better mechanical properties, easier
synthesis, and lower cost than single crystal SnSe [65]. However, most of ZT values of polycrystalline SnSe are much lower than that
of single crystals due to the low σ, low S and high κ [65]. To enhance S, arc-melting method was used to fabricate SnSe without
sintering, realizing a high S of ∼660 μV K−1 at 395 K, as well as an ultralow κ of 0.1 ∼0.2 W m−1 K−1 at the same temperature [88].
However, a low n of 7.95 × 1015 cm−3 resulted in a low σ of ∼0.16 S cm−1 at 395 K [88], which in turn resulted in a low S2σ, as well
as a low ZT. SnSe nanocrystals synthesized via the solvothermal route shows a n of ∼1 × 1019 cm−3 [118]. However, after sintering,
a low S of ∼160 μV K−1, as well as a high κ of ∼1.4 W m−1 K−1 were obtained at 300 K, resulting in a low ZT. Therefore, it is still a
challenge to decoupling σ, S, S2σ, and κ to realize high ZT for undoped polycrystalline SnSe.
As discussed in Section 4.3, single crystal SnSe has a peak ZT value of ∼2.6 along the b-axis [61]. In Section 5.2, the sintered
polycrystalline SnSe shows high anisotropy [66]. Extensive experiments [65,66,91,249] showed that the anisotropy became stronger
when texturing was applied in the undoped polycrystalline SnSe, leading to a better thermoelectric performance measured along the
⊥ direction than that measured along the // direction. This is because that along the ⊥ direction, most of the grains were arrayed
along the b- or c-axes [65,66,91,249], resulting in higher σ and higher S2σ.

5.3.1. Cold-pressing and hot-pressing


It has been found that the thermoelectric performances obtained from different texturing methods are different. For the cold-
pressing route, thermal annealing can influence both the microstructure and the thermoelectric properties via re-alignment of
crystalline domains [250], which has increased both S and σ. However, it was hard to achieve high ZT via cold-pressing due to the
high electrical resistivity [250], which derived from low σ [250]. Combined cold-pressing with annealing [250], a maximum ZT of
∼0.1 was achieved, which was still very low.
In the case of hot-pressing, through the combination of traditional melting, a highly textured polycrystalline SnSe ingot was
obtained, and a maximum ZT of 1.1 was obtained caused by a obtained high σ of ∼61.9 S cm−1 and a high S of ∼366.3 μV K−1, and
resulting in a high S2σ of ∼8.3 μW cm−1 K−2 at 873 K [66]. However, κ (∼0.66 W m−1 K−1) was very high in this case [66]. Strong
anisotropy was found in the sintered ingot due to the high-pressure and long-time sintering process. For the case of fusion method
combined with ball milling route [252], SnSe nanoparticles were sintered into ingot via hot-pressing, and the effect of hot-pressing
temperatures on the thermoelectric properties was explored. Due to the n optimization, the peak ZT value reached 0.73 for samples

317
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 27. (a) High resolution HAADF-STEM images of SnSe viewed along (a) a [1 0 0], (b) b [0 1 0], and (c) c [0 0 1] with overlays showing Sn atoms in
red and Se atoms in blue, (d) atomic-scale EDS mapping from zone axis [0 1 0], reproduced with permission [270], Copyright 2017, ACS Publications,
(e) EBSD phase map of SnSe with grains identified as Fm3m cubic rock-salt phase in red and Pnma orthorhombic phases in blue, (f) TEM image
showing rock-salt structured SnSe with inserted SAED pattern measured on the grain, reproduced with permission [141], Copyright 2017, RSC
Publishing, (g) HRTEM images of sintered SnSe pellet, reproduced with permission [119], Copyright 2016, Elsevier Publishing, (h) HRTEM images
of stripes with inserted corresponding inverse fast Fourier transform (IFFT) images of the dashed red rectangles, reproduced with permission [275],
Copyright 2017, John Wiley and Sons, (i) TEM image of sintered SnSe with preferred orientation along (4 0 0) and nanophase SnTe, reproduced with
permission [276], Copyright 2016, Elsevier Publishing.

hot-pressed at 673 and 723 K, indicating that the thermoelectric performance can be significantly improved by hot-pressing.

5.3.2. Melting with SPS


Although hot-pressing can secure a good quality of sintered bulks, the long-time sintering by hot-pressing can be time-consuming,
as shown in Table 6. To solve this problem, SPS was chosen. For SPS, the heat generation is internal during the sintering process, and
a high heating or cooling rate can be realized, contributing to a high density at a low sinter temperature. As a result, the SPS process is
fast, often within a few minutes, to both increase the efficiency of sintering and prevent the growth of grains. Thus, SPS has been
widely used in recent studies based on polycrystalline SnSe thermoelectric materials, and the thermoelectric performances were
comparable to the hot-pressing one (ZT = ∼1.0).
For the traditional melting and SPS route [65], anisotropic transport properties were observed, and a peak ZT of ∼0.5 was
obtained due to the moderate S2σ (∼4.0 μW cm−1 K−2) at 823 K [65]. This value was still higher than that via the combination of the
cold-pressing and annealing route [250]. However, κ (∼0.63 W m−1 K−1) was relatively high [65], which need further optimization.
When combining the melting, annealing, and SPS route [249], a pure phase with preferred orientation along the a -axis was obtained,
resulting in strong anisotropic thermoelectric properties. Although a low κ (∼0.4 W m−1 K−1 at 790 K) [249] was obtained via
annealing, the low n of 1.1 × 1017 cm−3 resulted in a low σ of ∼18.7 S cm−1 and a low S2σ of ∼2.7 μW cm−1 K−2 at 790 K, thus a
low peak ZT of ∼0.5 was obtained at 790 K. By combining the zone-melting with SPS route [89], highly textured structures and

318
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

strong anisotropic thermoelectric performance were obtained with a peak ZT of ∼1.0 along the // direction. This derived from a high
σ of ∼49.9 S cm−1, a high S of ∼408 μV K−1 and a very high κ of ∼0.7 W m−1 K−1 at 873 K [89].

5.3.3. Solvothermal with SPS


For traditional synthesis routes such as melting and mechanical alloying, it was a challenge to achieve high anisotropy in sintered
SnSe bulks due to the random orientated synthesized products before sintering. In order to achieve a high anisotropy in the syn-
thesized products, hydro- or solvothermal methods were used. Because the synthesized products via these routes are micro/nano-
sized single crystals, the products should possess strong anisotropy just like macro-sized single crystals. During the sintering process,
these micro/nano-sized single crystals can be sintered together with a very high anisotropy.
In the case of combining solvo-thermal with SPS route [81], the pure phase SnSe single crystals with a preferred orientation along
the (1 0 0) surfaces can be obtained, with an average size of less than 5 μm. This led to a high anisotropy of the corresponding bulks,
in which σ was greatly improved in the entire temperature range because of the increased n up to ∼1 × 1019 cm−3 [81]. An
enhanced S2σ was obtained and κ was lower than 1 W m−1 K−1 [81]. As a result, higher ZT were achieved in a relatively wide mid-
temperature range, which reached 0.44 at 522 K and 0.5 at 573 K [81]. These averaged ZT values are higher than these reported
averaged ZT values for single crystals [61]. To further increase the anisotropy and improve n, through combining the solvo-thermal
with SPS route [91], a very high peak ZT of ∼1.36 at 823 K was achieved [91]. The synthesized SnSe crystals (Sn0.98Se) have a
preferred orientation along (1 0 0) surfaces with a much larger average size of >100 μm [91]. As a result, the anisotropy of syn-
thesized products is much stronger. In addition, Sn vacancies of the synthesized SnSe crystals (Sn0.98Se) resulted in a large enhanced n
of ∼1.5 × 1019 cm−3 at 823 K, which in turn led to a high σ of ∼72.4 S cm−1, a high S of ∼309.9 μV K−1 and a low κ of
∼0.42 W m−1 K−1 at 823 K [91]. Consequently, a very high peak ZT of ∼1.36 was achieved [91]. Alternatively, Se vacancies can
lead to n-type polycrystalline SnSe1-x, but the thermoelectric performance was quite low [86].
Fig. 28 summarizes the recent obtained temperature-dependent thermoelectric properties (σ, S, S2σ, κ, and ZT) for undoped
polycrystalline SnSe prepared by different texturing routes discussed above. All properties were measured along the ⊥ direction. As
can be seen, all textured un-doped polycrystalline SnSe presented p-type thermoelectric properties. Fig. 28(f) presents the comparison
of peak and average ZT values of the obtained polycrystalline SnSe fabricated by different techniques, in which a peak ZT = 1.36 and
an average ZT = 0.68 were achieved through the combination of the solvo-thermal with SPS route [91], both represent the highest
values. Table 7 summarizes the comparisons of thermoelectric performance for undoped polycrystalline SnSe until now. As can be
seen, most of the ZT values are low than 1.5, indicating that new methodologies are highly needed to be developed.

5.4. Doping and alloying

In order to decouple σ, S, S2σ, and κ to realize a high ZT value for polycrystalline SnSe, doping is a useful method to tune n
[119,128,258,267–269,274,277–279] via controlling vacancies [280]. Besides, doping can also transfer the polycrystalline SnSe from
p-type to n-type [259,281,282]. For p-type doping, the dopants used include alkali metals (Li [283], Na [258,270,274,283–286], and
K [67,270,283]), II-A group alkali earth metals (Ca [127], Sr [127], Ba [127]), I-B group metals (Cu [127,269,287], Ag
[268,275–277,288]), II-B group metals (Zn) [127,262], III-A group metals (Al [269], In [269,279], Tl [278]), IV-A group elements
(Ge [127,289–291], Pb [269,286]), VI-A group metalloid (Te) [128,232,267], and lanthanides (Sm [292] and LaCl3 [293]). For n-
type doping, the dopants include halogens (Cl [235,281], Br [294], and I [259]), V-A group metals (Bi) [281]. Sometimes multiple
dopants were used, such as p-type Sn0.99Na0.01Se0.84Te0.16 [128] and n-type Sn0.74Pb0.20Ti0.06Se [295]. Most of the dopants were used
to substitute Sn sites to tune n, while Te was used to substitute Se sites [128,267].

5.4.1. Alkali
Because each alkali (such as Li, Na and K) atom has 1 valence electron, and in SnSe, each Sn atom has 2 valence electrons, alkali
metals are used to achieve p-type doping. Considering the alkali metals possess high activity, it is a challenge to dope them into SnSe
via solution routes such as hydro- or solvothermal method. Thus, melting routes are frequently used to realize alkali metal doping.
Practically, Na-doping is the most frequently used for alkali doping. As shown in Table 6, n was increased to ∼2.7 × 1019 cm−3 in
2 at.% Na-doped SnSe samples prepared by melting and hot-pressing [258]. Along the // direction, a peak ZT of ∼0.8 was achieved
via doped with 1.5% Na due to the obtained low κ of ∼0.33 W m−1 K−1 at 773 K [258]. Such a low κ value derived from the induced
point defects. For K-doping, K0.01Sn0.99Se polycrystals were prepared by combining mechanical alloying and SPS [67]. The maximum
S2σ was enhanced significantly from ∼2.8 μW cm−1 K−2 for undoped SnSe to ∼3.5 μW cm−1 K−2 for K-doped SnSe [67]. The
absence of Sn oxides at grain boundaries and nanoprecipitates contributed to an ultralow κ of ∼0.24 W m−1 K−1, resulting in a peak
ZT of ∼1.1 via doping with 1% K at 773 K along the ⊥ direction [67]. In the case of Na and K co-doping [270], the energy offsets of
multiple valence bands in SnSe were decreased via Na-doping and further reduced by (Na, K) co-doping, resulting in an enhancement
in S and an increase in S2σ to 4.92 μW cm−1 K−2. In this case, κl of polycrystalline SnSe was decreased by the introduction of effective
phonon scattering centres such as point defects and anti-phase boundaries, which decreased to as low as 0.29 W m−1 K−1 at 773 K
[270]. Thus, a high peak ZT of 1.2 was obtained for 1% (Na, K) co-doped SnSe [270]. In another Li-doped case, S2σ of ∼4 μW cm−1
K−2, κ of ∼0.7 W m−1 K−1 and peak ZT of ∼0.5 were obtained at 800 K through combining melting and SPS [283].

5.4.2. I-B and II-B group metal


Similar to alkali metals, I-B group metals (such as Cu and Ag) have 1 valence electron in each atom, so that I-B group metals are
also good candidates to realize the improvement of thermoelectric performance by p-type doping. Different from alkali metals, I-B

319
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 160 M + HP b 800 M + HP M + A + SPS


140 M + A + SPS SS + SPS ZM + SPS
SS + SPS 700 M + SPS CP + A
120
ZM + SPS

S (μV K-1)
σ (S cm-1)
100 600
M + SPS
80 CP + A 500
60 400
40
300
20
0 200
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
c 12 M + HP M + A + SPS d 1.6
M + HP
SS + SPS ZM + SPS M + A + SPS
10
S2σ (μW cm-1K-2)

M + SPS CP + A SS + SPS

κ (W m-1K-1)
8 1.2 ZM + SPS
M + SPS
6 CP + A
0.8
4

2 0.4
0
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
e 1.4 M + HP f 1.4 1.36 Average
M + A + SPS Peak
1.2 1.2
SS + SPS 1.1
1.04
1.0 ZM + SPS 1.0
M + SPS 0.8
0.8
ZT
ZT

CP + A 0.68
0.6 0.6 0.53
0.44
0.4 0.4
0.25 0.25
0.2 0.16 0.19
0.2 0.13
0.06
0.0 0.0
300 400 500 600 700 800 900 M M+A SS ZM M CP
+ + + + + +
T (K) HP SPS SPS SPS SPS A

Fig. 28. Temperature-dependent thermoelectric properties for pure polycrystalline SnSe prepared by melting + hot-pressing [66], melting + an-
nealing + SPS [249], solve-thermal + SPS [91], zone melting (ZM) + SPS [89], melting + SPS [65], and cold-pressing + annealing [250]: (a) σ, (b)
S, (c) S2σ, (d) κ and (e) ZT. (f) Comparison of both average and peak ZT of different materials from 323 K to 873 K.

group metals such as Cu have been demonstrated to be able to dope into SnSe system via solution routes [296]. Thus, there are more
choices for the doping of I-B group metals during synthesis process. Besides, other metals such as II-B group metals (Zn) were used to
realize the p-type doping.
For the case of Cu doping, by employing conventional fusion method followed by SPS, 2% Cu doped SnSe showed an enhanced
peak ZT of ∼0.7 ± 0.02 at 773 K [269]. This enhancement primarily derived from the presence of Cu2Se second phase associated
with intrinsic SnSe nanostructures. A high n of 1.84 × 1020 cm−3 was achieved via Cu-doping, contributing a moderate σ of ∼20 S
cm−1 and a moderate S of ∼250 μV K−1 at 773 K [269]. However, κ was very low (∼0.3 W m−1 K−1 at 773 K) [269], thus a good
peak ZT was achieved. In another case of doping with Cu, Sn1-xCuxSe (x = 0–0.03) was obtained by combining melting with high
pressure (6.0 GPa) sintering, in which maximum S2σ and ZT values were 3.78 μW cm−1 K−2 and 0.79 for Sn0.97Cu0.03Se at 823 K, and
the obtained κ was relatively low (∼0.4 W m−1 K−1) (HPS) [287].
For the case of Ag doping, by employing conventional melting method followed by SPS [277], a peak ZT of 0.74 was achieved via
doping with 1% Ag at 823 K [277], in which the doped Ag can tune n to a high value of 1.9 × 1019 cm−3 at room temperature,
resulting in a high S2σ of ∼6.0 μW cm−1 K−2 at 823 K. For another case of Ag doping, undoped SnSe polycrystals with novel
nanolaminar structures have been fabricated via conventional solid-state reaction at high temperature [275]. The phase transition
contributes to the enhancement of n and phonon scattering above 600 K, and the nanolaminar polycrystals significantly improved the

320
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Table 7
A comprehensive summary on the thermoelectric performance of undoped polycrystalline SnSe. The marked (r) in n column means the n values
were tested at room temperature.
Product Type Direction Synthetic ZT T (K) σ (S cm−1) S (μV K−1) S2σ (μW κ (W ρ (g cm−3) n (cm−3) Year Ref.
Method cm−1 K−2) m−1
K−1)

SnSe p ⊥ M + SPS ∼0.5 823 ∼31.6 ∼355.7 ∼4.0 ∼0.63 5.93 – 2014 [65]
SnSe p – AM – 395 – ∼660 – 0.1 ∼0.2 – 7.95 × 1015 (r) 2015 [88]
SnSe p – ST + SPS – 300 – ∼160 – ∼1.4 ∼5.99 ∼1 × 1019 2015 [118]
SnSe p ⊥ M+A + ∼0.5 790 ∼18.7 ∼379.8 ∼2.7 ∼0.40 5.92 1.1 × 1017 (r) 2015 [249]
SPS
SnSe p – M + SPS – 850 ∼66.0 ∼300 ∼6.0 – 6.13 – 2016 [90]
SnSe p ⊥ M + HP 1.1 873 ∼61.9 ∼366.3 ∼8.3 ∼0.66 6.05 4.0 × 1017 (r) 2016 [66]
SnSe p ⊥ M + SPS ∼1.0 873 ∼49.9 ∼408.0 ∼8.3 ∼0.70 6.03 5.25 × 1017 (r) 2016 [89]
SnSe p ⊥ ST + SPS ∼0.6 773 ∼28.7 ∼339.8 ∼3.3 ∼0.44 5.87 1.02 × 1019 (r) 2016 [81]
SnSe p ⊥ CP + A ∼0.1 772 ∼15.3 ∼279.8 ∼1.2 ∼0.75 – 2016 [250]
SnSe p ⊥ M + HP 0.73 800 ∼64.3 ∼322.7 ∼6.7 ∼0.73 – 5.57 × 1016 2017 [252]
SnSe p ⊥ HT + CP + 0.54 550 ∼13.9 360 ∼1.8 ∼0.18 – ∼2.5 × 1018 (r) 2017 [423]
HP
SnSe p ⊥ MA + SPS ∼0.71 873 ∼43.2 ∼292.8 3.7 ∼0.45 6.14 4.18 × 1017 (r) 2017 [424]
SnSe p – AM – 625 – 1050 – – – – 2017 [422]
SnSe p ⊥ ST + HP ∼0.83 803 ∼16.7 ∼346.0 ∼2.0 ∼0.2 – ∼1.75 × 1019 2017 [233]
SnSe p ⊥ CM + SPS 0.5 773 ∼25.0 ∼368.8 ∼3.4 ∼0.53 ∼5.99 3.4 × 1017 (r) 2018 [150]
Sn0.98Se p ⊥ ST + SPS 1.36 823 72.4 309.9 6.9 0.42 6.09 1.48 × 1019 2018 [91]
SnSe1−x n ⊥ ST – 400 ∼0.4 ∼−130.0 ∼0.007 ∼0.24 – 2016 [86]
SnSe0.98 n // M + HP 0.07 773 ∼4.5 ∼−364.3 ∼0.6 ∼0.47 6.05 – 2018 [282]

performance via promoting μ and simultaneously strengthening the phonon scattering [275]. 1.5% Ag-doped nanolaminar SnSe
polycrystals realized in a high S2σ of ∼6.0 μW cm−1 K−2, an extremely low κl of 0.23 W m−1 K−1, and a high peak ZT of 1.3 at 773 K
[275].
For the case of doping with II-B group metals (such as Zn) [262], through combining traditional melting with hot-pressing route,
the obtained Zn0.01Sn0.99Se polycrystals showed a peak ZT of ∼0.96 at 823 K [262]. Zn dopants can decouple σ and S to achieve a
high S2σ of ∼8.0 μW cm−1 K−2 via tuning n to an appropriate value of ∼4.5 × 1018 cm−3.

5.4.3. Halogen
Because the valence state of halogens (such as Cl, Br and I) is −1 in most compounds, and Se in SnSe always present −2, halogens
are good candidates to tune n for p-type doping, as well as realize the n-type doping. Considering halogens possess high reactivity,
similar to the alkali metals, they are difficult to be doped into SnSe via solution routes such as hydro- or solvo-thermal method. Thus,
melting routes are frequently used to realize halogen doping.
For realizing p-type Cl-doping, solid-state reaction was used to fabricate SnSe0.985Cl0.015 polycrystals using precursors of Sn, Se
and SnCl2 [275]. From which, a high S2σ of ∼4.5 μW cm−1 K−2 and a low κ of ∼0.23 W m−1 K−1 were achieved, leading to a high
peak ZT of 1.1 at 773 K [275]. On the other hand, n-type Cl doping can be realized from the aqueous solution method for the facile
synthesis of Cl-doped SnSe nanoparticles [235]. After sintering by hot-pressing, a low σ of ∼4 S cm−1 and a moderate S of
∼−300 μV K−1 were obtained, resulting in a low S2σ of ∼0.02 μW cm−1 K−2 at 540 K. Although such thermoelectric performances
are low, this study provided a simple, rapid, and low-cost method to fabricate n-type SnSe thermoelectric materials. In the case of n-
type BaCl3-doping, SnSe0.95-BiCl3 compounds were synthesized through melting with SPS, in which a peak ZT of ∼0.7 was obtained
via doping with 0.4% BiCl3 at 793 K [281]. In this case, doped BiCl3 can decouple σ and S to achieve a high S2σ of ∼5.0 μW cm−1
K−2.
Except Cl, Br and I were also used as n-type dopants. PbBr2-doped polycrystalline SnSe has been synthesized using the combi-
nation of melting with hot-pressing and has an improved n of 1.86 × 1019 cm−3 [294]. Consequently, an increased σ of 40 ± 2 S
cm−1 and a high S2σ of ∼4.8 μW cm−1 K−2 were achieved at 713 K [294]. Since κ of ∼0.7 W m−1 K−1 was relatively high, a
moderate peak ZT of 0.54 ± 0.1 was obtained at 793 K along the ⊥ direction. By using melting and hot pressing, n-type I-doped SnSe
polycrystals can be obtained [259]. The 3% I doped SnSe has a peak ZT of ∼1.0 due to the obtained high S2σ of ∼3.9 μW cm−1 K−2
and a low κ of ∼0.3 W m−1 K−1) at 773 K [259].

5.4.4. IV-A and VI-A group


Sn belongs to the IV-A group, and Se belongs to the VI-A group, so that IV-A group elements (such as Ge and Pb) can substitute Sn
atoms, and VI-A group elements (such as S and Te) can substitute Se atoms during the doping. These dopants can induce point defects,
leading to further improvement of the thermoelectric performance of SnSe.
For example, Ge-doped nanostructured SnSe has been prepared by an arc-melting method [289]. This Ge-doped nanostructured
SnSe shows a semiconductor behavior with resistivity values higher than that of the pure SnSe compound, because Ge does not
donate carriers [289]. However, S has been improved significantly and reaches to a value of 1000 μV K−1 with a low Ge-doping level
[289]. Due to a high density of point defects caused by Ge-doping, the obtained κ was extremely low to be ∼0.1 W m−1 K−1 at 385 K.

321
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 140 No doping 1.5 % Ag b 900 No doping 1 % Na, K


1 % Na, K 1 % Te 1 % Ge 0.4 % BiCl3
120
1 % Ge 1 % Zn 600
100 0.4 % BiCl3 3%I

cm-1)

S ( V K-1)
80 300
60 1.5 % Ag 1 % Te
(S
0 1 % Zn 3%I
40
20 -300
0
-600
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
c 12 No doping 1.5 % Ag d 2.0 No doping 1.5 % Ag
1 % Na, K 1 % Te 1 % Na, K 1 % Te
W cm-1K-2)

10 1 % Ge 1 % Zn 1 % Ge 1 % Zn
1.6

m-1K-1)
8 0.4 % BiCl3 3%I 0.4 % BiCl3 3%I

6 1.2

(W
4 0.8
(

2
S2

0.4
0
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
e 1.6 No doping 1.5 % Ag f 1.3 Average
1.4 1 % Na, K 1 % Te 1.2 1.2
1.1 Peak
1.2 1 % Ge
1 % Zn 0.96
1.0 0.9 0.84
0.4 % BiCl3
ZT

0.8
ZT

3%I 0.61 0.6


0.6 0.6 0.530.5 0.55 0.57

0.4
0.2 0.3 0.26
0.16 0.15 0.19 0.18
0.0
0.0
300 400 500 600 700 800 900 Pure Ag Na, K Te Ge Zn BiCl3 I
T (K) 1.5 % 1 % 1 % 1 % 1 % 0.4 % 3 %

Fig. 29. Temperature-dependent thermoelectric properties for pure [65] and doped with 1.5% Ag [275], 1% (Na, K) [270], 1% Te [232], 1% Ge
[290], 1% Zn [262], 0.4% BiCl3 [281] and 3% I [259] polycrystalline SnSe: (a) σ, (b) S, (c) S2σ, (d) κ and (e) ZT. (f) Comparison of both average and
peak ZT of different materials from 323 K to 873 K.

For another case of Ge-doping, polycrystalline Sn1-xGexSe were fabricated using a solid solution route [290]. The doping of Ge has a
mild effect on the κ reduction (∼0.5 W m−1 K−1 at 800 K for Sn0.92Ge0.08Se), while S2σ was significantly deteriorated (1.87 μW cm−1
K−2 at 823 K for the Sn0.92Ge0.08Se), leading to a peak ZT of 0.6 for Sn0.96Ge0.04Se at 823 K. On the contrast, in the other Ge-doped
case [291], it was found that Ge is not a suitable dopant because the peak ZT of Sn0.97Ge0.03Se was only 0.15 at 773 K although a
decreased κ of ∼0.25 W m−1 K−1 at 773 K was observed.
In terms of VI-A group doping, Te has been widely used. For an example, SnSe0.98Te0.02 has been fabricated by a simple and rapid
high pressure and high temperature (HPHT) method [119]. By narrowing the bandgap of SnSe to 0.746 eV, a moderate S of
∼330 μV K−1, as well as a maximum S2σ of ∼1 μW cm−1 K−2 were obtained at 3 GPa. For other case of n-type Te-doping,
SnSe0.9375Te0.0625 bulks were fabricated by melting method [267]. The bandgap of SnSe reduced from 0.643 to 0.608 eV after Te-
doping [267], but a low σ of ∼3.5 S cm−1 and a moderate S of ∼−270 μV K−1 led to a low S2σ of ∼0.7 μW cm−1 K−2 at 673 K.
Recently, a facile microwave-assisted solvo-thermal method has been used to fabricate large-scale SnSe1-xTex nanoplates [232]. After
sintering via SPS, both peak ZT of 1.1 at 800 K and an average ZT of 0.56 from 300 to 800 K were achieved in the p-type SnSe0.9Te0.1
pellet. The reduced κ of ∼0.52 W m−1 K−1 at 800 K was attributed to the enhanced phonon scattering by point defects, and the
enhanced S2σ of ∼6.8 μW cm−1 K−2 was caused by the secured high preferential orientation and the increased n of 1 × 1019 cm−3
[232].

322
Z.-G. Chen et al.

Table 8
A comprehensive summary on the thermoelectric performance of doped polycrystalline SnSe. The marked (r) in n column means the n values were tested at room temperature.
Product Type Direction Synthetic Method ZT T (K) σ (S cm−1) S (μV K−1) S2σ (μW cm−1 K−2) κ (W m−1 K−1) n (cm−3) ρ (g cm−3) Year Ref.

Ag0.01Sn0.99Se p // M + A + HP 0.6 750 ∼45.9 ∼344.1 ∼5.4 ∼0.68 ∼3.5 × 1018 (r) ∼5.93 2014 [268]
Sn0.99Na0.01Se0.84Te0.16 p ⊥ MA + SPS 0.72 773 ∼67.4 ∼275.0 ∼5.1 ∼0.50 – – 2015 [128]
Na0.01Sn0.99Se p ⊥ M + SPS 0.75 823 ∼49.6 ∼311.1 4.8 ∼0.53 1.0 × 1019 (r) ∼5.99 2016 [274]
Na0.015Sn0.985Se p // M + MA + HP ∼0.8 773 ∼37.9 ∼298.8 ∼3.4 ∼0.33 ∼2.1 × 1019 (r) 5.81 2016 [258]
Na0.01Sn0.99Se p ⊥ M + SPS ∼0.8 800 ∼81.2 ∼267.2 ∼5.8 ∼0.50 ∼1.5 × 1019 – 2016 [283]
Na0.02Sn0.98Se p ⊥ High pressure + SPS 0.87 798 ∼56.4 ∼288.8 4.7 0.4 3.08 × 1019 (r) ∼5.62–5.81 2017 [285]
Na0.005 K0.005Sn0.99Se p ⊥ MA + SPS 1.2 773 ∼34.9 ∼374.7 ∼4.9 0.32 ∼7.2 × 1019 (r) 5.71 2017 [270]
Na0.005Sn0.995SeCl0.005 p ⊥ SSR + HP 0.84 810 ∼79.2 ∼228.6 ∼4.1 ∼0.39 ∼3.95 × 1019 (r) ∼5.93 2017 [426]
Na0.01(Sn0.96Pb0.04)0.99Se p ⊥ M + SPS ∼1.2 773 ∼89.4 ∼269.7 ∼6.5 ∼0.45 ∼2.8 × 1019 – 2017 [286]
Ag0.01Sn0.99Se p ⊥ M + A + SPS 0.74 823 ∼54.8 ∼330.9 6.0 ∼0.66 1.9 × 1019 (r) ∼5.99 2016 [277]
Ag0.03Sn0.97Se p ⊥ ST + SPS 0.8 850 ∼90.3 ∼266.2 ∼6.4 ∼0.68 9 × 1018 (r) >5.56 2017 [288]
Sn0.995Tl0.005Se p ⊥ M + HP 0.6 725 ∼68.9 ∼300.0 ∼6.2 ∼0.75 – ∼5.99 2016 [278]
Sn0.98Cu0.02Se p – M + A + SPS 0.7 773 ∼42.4 ∼238.6 ∼2.4 0.27 1.84 × 1020 (r) ∼6.12 2016 [269]
Sn0.97Cu0.03Se p – M + HP 0.79 823 ∼35.0 ∼325.1 ∼3.7 ∼0.39 1.57 × 1017 (r) 6.16 2017 [287]
Sn0.99Cu0.01Se p – HT + SPS 1.2 873 ∼35.2 ∼310.6 ∼3.4 0.2 – – 2018 [296]

323
Sn0.99In0.01Se p // M + HP 0.2 823 ∼6.53 ∼350.0 ∼0.8 ∼0.36 ∼2.9 × 1017 (r) ∼5.87 2015 [279]
Sn0.9Ge0.1Se p – M – 400 – ∼843.2 – ∼0.39 – – 2016 [289]
Sn0.96Ge0.04Se p ⊥ ZM + HP 0.6 823 35.6 ∼378.5 5.1 ∼0.7 3.36 × 1017 (r) >5.81 2017 [290]
K0.01Sn0.99Se p ⊥ MA + SPS ∼1.1 773 ∼18.6 ∼421.4 ∼3.3 ∼0.24 9.2 × 1018 (r) – 2016 [67]
Zn0.01Sn0.99Se p ⊥ M + HP 0.96 873 ∼74.1 ∼328.5 8.0 ∼0.73 ∼4.5 × 1018 – 2016 [262]
Na0.01Sn0.99Se p ⊥ M + A + SPS 0.85 800 ∼100.4 ∼271.5 ∼7.4 ∼0.50 ∼6.5 × 1019 (r) 5.94 2017 [284]
Ag0.015Sn0.985Se p – M 1.3 773 ∼44.7 ∼344.0 ∼5.2 ∼0.30 ∼8 × 1018 (r) 5.87 2017 [275]
SnSe0.985Cl0.015 p – M 1.1 773 ∼25.5 ∼399.3 ∼4.1 ∼0.30 ∼1 × 1017 (r) 5.87 2017 [275]
Sn0.97Na0.03Se p ⊥ SPS 0.82 773 ∼65.1 ∼280.2 ∼5.1 ∼0.50 ∼2.2 × 1019 ∼5.93 2017 [425]
SnSe0.9Te0.1 p ⊥ ST + SPS 1.1 800 ∼57.4 ∼322.8 ∼6.0 ∼0.44 ∼1 × 1019 (r) ∼5.87 2017 [232]
Sn0.97Sm0.03Se p – M + HP 0.55 823 ∼33.6 ∼250.0 ∼2.1 ∼0.32 ∼1.27 × 1017 (r) – 2017 [292]
SnSe-1.0 wt% LaCl3 p ⊥ MA + SPS 0.55 750 ∼15.6 ∼350.6 ∼1.9 ∼0.27 1.53 × 1016 (r) ∼5.93 2018 [293]
SnSe0.9375Te0.0625 n ⊥ M + HP – 673 ∼3.63 ∼−276.2 ∼0.3 – −2.1 × 1017 (r) ∼5.87 2012 [267]
SnSe0.95–0.4% BiCl3 n ⊥ M + SPS 0.7 793 ∼28.9 ∼−414.0 ∼5.0 ∼0.60 −1.07 × 1019 (r) – 2016 [281]
SnSe0.95–3% PbBr2 n ⊥ M + HP 0.54 793 ∼36.0 ∼−360.0 ∼4.7 ∼0.72 −1.86 × 1019 (r) 5.99 2017 [294]
SnSe0.87S0.1I0.03 n // M + MA + HP ∼1.0 773 ∼10.0 ∼−624.5 ∼3.9 ∼0.30 −3.8 × 1016 (r) 5.75 2016 [259]
SnSe-Cl (0.6%) n ⊥ HT + HP – 540 ∼9.0 ∼−264.8 ∼0.6 – −6.43 × 1018 (r) ∼5.87 2017 [235]
Sn0.74Pb0.20Ti0.06Se n ⊥ MA + SPS 0.4 773 ∼14.8 −450 3.0 ∼0.58 2.47 × 1016 (r) 2017 [295]
Progress in Materials Science 97 (2018) 283–346
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

To summarize, Fig. 29(a)–(e) shows the main temperature-dependent thermoelectric properties (σ, S, S2σ, κ, and ZT) for undoped
[65] and doped with 1.5% Ag [275], 1% (Na, K) [270], 1% Te [232], 1% Ge [290], 1% Zn [262], 0.4% BiCl3 [281] and 3% I [259]
polycrystalline SnSe, and the more detailed thermoelectric performances of doped polycrystalline SnSe are summarized in Table 8. As
can be seen, most of the dopants can enhance σ in a wide temperature range by tuning n, resulting an improvement in S2σ. It should
be noted that compared with undoped polycrystalline SnSe, most of the dopants can further decrease κ in a wide temperature range
(between 323 K and 773 K).
In fact, the induced dopants can achieve large quantities of point defects and/or nanoprecipitates, which cause strain fields
around substituted Sn or Se atoms and/or nanoprecipitates. In this respect, phonon scattering effects can be strengthened, leading to
decreasing κl [61]. Fig. 29(f) presents a comparison of peak and average ZTs for polycrystalline SnSe doped with different dopants,
and shows that a peak ZT of 0.98 and an average ZT of 0.49 via doped with 1% K [67] both represent the highest values. These results
indicate that doping is a promising method to enhance the thermoelectric performance of polycrystalline SnSe, as well as a con-
venient way to realize n-type thermoelectric materials with a high thermoelectric performance for polycrystalline SnSe.

5.5. Inclusions

The introduction of appropriate inclusions in polycrystalline SnSe can achieve a high peak ZT in SnSe system [84,276]. Many
different types of inclusion, such as second phase (SnSe2 [84] and Se nanowires [297]), nanoprecipitates (rock-salt-type SnSe [141]
and Sn oxides [67]), simple substance (carbon black nanoinclusions [298]), compounds (SnS [129,299], 2D-MoSe2 [300]) and
compounds (SiC [259] and PbTe [301]) have been developed in SnSe. Sometimes multiple inclusions, such as (MoS2 + graphene)
also developed [85]. In some case, both dopant and inclusion were induced at the same time, such as p-type SnTe/Ag0.005Sn0.995Se
composites which was SnSe doped with 0.5% Ag and incorporated with 0.5 mol% SnTe [276]; Ag0.01Sn0.99Se1−xSx compounds which
was SnSe doped with 1% Ag and incorporated with SnS [302]; and n-type (Sn0.9Se0.87I0.03)(SnS)0.1 which was 90% SnSe doped with
3% I and alloyed with 10% SnS [257]. SnSe can also be treated as inclusion in other thermoelectric materials systems, such as
(Cu2Se)0.97(SnSe)0.03 [303], and (PbSe)0.9(SnSe)0.1 [257].
It is difficult to compare the thermoelectric performance of inclusion/SnSe composites because the induced inclusions show
different mechanism. In this respect, we only briefly summarize some typical cases here.

5.5.1. Solid solution and co-synthesis


Due to the extremely low κ, solid solution, such as SnS–SnSe [129], PbSe–SnSe [284] and SnTe–SnSe [304] system, shows a
potential way to optimize thermoelectric performance [129]. SnS shows analogous band structure and electrical properties with
SnSe, making SnSe to be a good candidate to form solid solution such as SnS–SnSe system. In SnSe–SnS solid-solution, a low κ of
∼0.3 W m−1 K−1, a high S of ∼600 μV K−1, and a peak ZT of 0.82 was achieved for SnS0.2Se0.8 at 823 K via melting route [129].
Besides, the solid solution system also exhibited anisotropic thermoelectric performance along the // and ⊥ directions [129]. Other
case showed that by mechanical alloying followed by SPS, SnS1-xSex solid solutions also exhibited anisotropic thermoelectric per-
formance [299]. The κ decreased with increasing the Se content and fell to 0.36 W m−1 K−1 at 823 K for the composition SnS0.5Se0.5
[299]. The optimized S2σ of ∼3.7 μW cm−1 K−2 and the reduced κ of ∼0.5 W m−1 K−1 resulted in a peak ZT value of 0.64 at 823 K
for SnS0.2Se0.8 along the // direction [299]. Except for the SnSe–SnS solid-solution system, single-phase SnSe–PbSe solid-solution
(Sn1-xPbxSe) was obtained via alloying [284]. κl was reduced from 1.4 to 0.85 W m−1 K−1 by 12 at.% PbSe alloying, derived from the
strain and mass fluctuations [284]. Interestingly, S and n were nearly unchanged by Pb substitution, indicating a constant m∗ and an
undisrupted VBM [284].
Except for solid-solution, the SnSe–PbSe system can be achieved via hydro-thermal route [84]. During hydro-thermal synthesis
process, SnSe and PbSe were synthesized together at the same synthesis conditions [84]. A maximum ZT of ∼1.7 was achieved for 1%
PbSe/SnSe composites [84] due to a high S2σ of ∼8.9 μW cm−1 K−2, as well as a low κ of ∼0.25 W m−1 K−1 at 873 K [84]. Such high
S2σ is derived from the induced PbSe phase and such low κ is attributed to the enhanced phonon scattering at all-scale hierarchical
architectures (from the nanoscale precipitates to mesoscale grains) [84].

5.5.2. Coupled with doping and multiple inclusions


To improve the thermoelectric performance of SnSe, both dopant and inclusion are induced in SnSe system and multiple in-
clusions have been developed [84,276,305,306]. For the case of inducing both Ag as dopant and SnTe as inclusion, a peak ZT of 1.6
with a high σ of ∼112.3 S cm−1 and a high S2σ of ∼11 μW cm−1 K−2 at 875 K was achieved for 0.5 mol% SnTe/Ag0.005Sn0.995Se
composites [276]. In Ag0.005Sn0.995Se fabricated via solution method [276], the 0.5% doped Ag resulted in n of ∼3.99 × 1018 cm−3
at room temperature [276], and the induced 0.5 mol% SnTe resulted in enhanced S2σ of ∼11 μW cm−1 K−2 and reduced κl of
∼0.4 W m−1 K−1 [276] at 875 K. Furthermore, these polycrystalline composites have high peak ZT (> 1) from 800 to 900 K and
good experimental repeatability [276]. For the SnSe-PbSe solid-solution with Na as dopant [84], a maximum ZT of ∼0.7 was
achieved for Sn0.88Pb0.12Se–1% Na [276] due to a high n of ∼6.5 × 1019 cm−3 adjusted via the doping of Na and relative a low κ of
∼0.5 W m−1 K−1 derived from the solid-solution.
For the case of multiple inclusions, by rapid induction melting followed by rapid hot-pressing, a peak ZT of 0.98 and a low κ of
∼0.39 W m−1 K−1 were achieved at 810 K for 3.2 wt% MoS2/graphene/SnSe composites [85]. Such low κ derived from the induced
layered MoS2 + graphene composites [85]. For the case of inducing 2D-materials (MoSe2) [300], by the solid solution method
followed by SPS, introducing 2D-MoSe2 into SnSe matrix achieved a high n of ∼2 × 1019 cm−3 without distinctly deteriorating κ
[300], leading to a peak ZT of 0.5 at 773 K [300].

324
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 150 Pure + SnS b 900 Pure + SnS


+ SnTe/Ag + PbSe + SnTe/Ag + PbSe
120 + MoS2/graphene 750 + MoS2/graphene

cm-1)

S ( V K-1)
+ MoSe2 + MoSe2
90 600

60
(S
450
30
300
0
150
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
c 14 Pure + SnS d Pure + SnS
1.6 + SnTe/Ag + PbSe
12 + SnTe/Ag + PbSe
W cm-1K-2)

+ MoS2/graphene + MoS2/graphene

m-1K-1)
10 1.2 + MoSe2
+ MoSe2
8
0.8

(W
6
(

4
S2

0.4
2
0
300 400 500 600 700 800 900 300 400 500 600 700 800 900
T (K) T (K)
e 1.8 Pure f 1.8 1.71 1.72

+ SnS 1.5
Average
1.5
+ SnTe/Ag Peak
1.2 + PbSe 1.2
+ MoS2/graphene 1.01
ZT
ZT

0.9 0.9
+ MoSe2 0.62
0.62 0.5
0.6 0.6 0.53 0.46
0.47

0.3 0.3 0.16 0.17 0.13

0.0 0.0 Pure + + + + +


300 400 500 600 700 800 900 SnS SnTe PbSe MoS2 MoS2
T (K) Ag graphene

Fig. 30. Temperature-dependent thermoelectric properties for pure SnSe [65], (SnSe)0.8(SnS)0.2 [129], 0.5 mol% SnTe + Ag0.005Sn0.995Se compo-
sites [276], 1% PbSe + SnSe composites [84], 3.2 wt% MoS2/graphene + SnSe composites [85], and 1.5% MoSe2 + SnSe composites [300]: (a) σ,
(b) S, (c) S2σ, (d) κ and (e) ZT. (f) Comparison of both average and peak ZT of different materials from 323 to 873 K.

To summarize, Fig. 30(a)–(e) plots temperature-dependent thermoelectric properties (σ, S, S2σ, κ, and ZT) for pure SnSe [65],
(SnSe)0.8(SnS)0.2 [129], 0.5 mol% SnTe + Ag0.005Sn0.995Se composites [276], 1% PbSe + SnSe composites [84], 3.2 wt% MoS2/
graphene + SnSe composites [85], and 1.5% MoSe2 + SnSe composites [300] discussed above. Fig. 30(f) also presents a comparison
of the peak and average ZTs for polycrystalline SnSe incorporating with different inclusions, and shows that a peak ZT of 1.72
presents the highest value via incorporating with 1% PbSe [84], and an average ZT = 0.62 presents the highest value via in-
corporating with 0.5% mol SnTe and doped with 0.5% Ag [276]. Table 9 also summarizes the thermoelectric performance of
polycrystalline SnSe incorporated with different inclusions in more detail.

6. Two-dimensional SnSe

The initial investigations on two-dimensional (2D) thermoelectric materials [307,308] predicted that the quantum confinement
effect of electronic carriers in low-dimensional materials can enhance S2σ [307,308]. In Section 6, we discuss the structural char-
acteristics of monolayer SnSe in detail, and summarize the recent advances in 2D SnSe-based thermoelectric materials, such as thin
film, nanosheets, and organic/inorganic composites.

325
Z.-G. Chen et al.

Table 9
A comprehensive summary on the thermoelectric performance of polycrystalline SnSe incorporated with other compounds. The marked (r) in n column means the n values were tested at room
temperature.
Product Type Direction Synthetic method ZT T (K) σ (S cm−1) S (μV K−1) S2σ (μW cm−1 K−2) κ (W m−1 K−1) n (cm−3) ρ (g cm−3) Year Ref.

20
(Cu2Se)0.97(SnSe)0.03 p – M + SPS 1.41 823 ∼302.4 ∼199.2 12.0 ∼0.70 9.93 × 10 (r) ∼6.61 2015 [303]
(SnS)0.2(SnSe)0.8 p // M + A + SPS 0.82 823 ∼13.8 ∼466.9 3.0 ∼0.30 ∼3.0 × 1017 (r) ∼5.93 2015 [129]
0.5 mol% SnTe/Ag0.005Sn0.995Se composites p // M + SPS 1.6 875 ∼112.3 ∼313.0 ∼11.0 0.5 3.99 × 1018 (r) ∼6.07 2016 [276]
1% PbSe/SnSe composites p // HT + SPS ∼1.7 873 ∼74.8 ∼344.9 ∼8.9 ∼0.55 6.1 × 1018 (r) – 2016 [84]
Sn oxide nanoprecipitates/K0.01Sn0.99Se p ⊥ MA + SPS ∼1.1 773 ∼18.6 ∼421.4 ∼3.3 ∼0.24 9.2 × 1018 (r) – 2016 [67]
(SnS)0.2(SnSe)0.8 p // MA + SPS 0.64 823 ∼26.3 ∼332.1 ∼2.9 ∼0.38 2.1 × 1017 (r) 5.45 2017 [299]
3.2 wt% MoS2/graphene/SnSe composites p ⊥ M + HP 0.98 810 ∼70.5 ∼258.2 ∼4.7 ∼0.39 2.49 × 1019 (r) ∼5.93 2017 [85]

326
1.5% MoSe2/SnSe composites p ⊥ M + SPS 0.5 773 ∼21.8 ∼428.3 ∼4.0 ∼0.64 ∼2 × 1019 – 2017 [300]
Rock-salt-type nanoprecipitates/SnSe p // HT + SPS 1.3 850 ∼48.6 ∼288.5 ∼4.0 ∼0.26 5.6 × 1018 (r) – 2017 [141]
SnSe2/Sn0.98Se p ⊥ M + A + SPS 0.61 848 ∼48.3 ∼308.6 ∼4.6 ∼0.64 – ∼5.99 2017 [84]
Ag0.01Sn0.99Se0.85S0.15 p // M + A + HP 1.67 823 ∼25 296.4 ∼2.2 0.11 7.6 × 1018 (r) – 2017 [302]
2.5 vol% SnSe/C composites p ⊥ M + HP 1.21 903 ∼100.2 ∼310.0 9.6 ∼0.72 >1.8 × 1019 ∼5.81–6.06 2016 [298]
1.5 vol% SnSe/PbTe composites p ⊥ M + HP 1.26 880 ∼94.7 ∼301.2 9.2 ∼0.64 ∼1.94 × 1019 – 2016 [301]
Na0.015Sn0.985Se/CNTs composites p ⊥ M + A + SPS ∼0.96 773 ∼55.6 ∼299.2 ∼5.0 ∼0.4 ∼4 × 1019 (r) ∼6.06 2018 [305]
Sn0.995Ag0.005S0.2Se0.8 p ⊥ MA + SPS ∼1.1 823 ∼53.4 ∼315.1 ∼5.3 ∼0.4 ∼2 × 1019 5.78, 5.68, 5.56 2018 [306]
SiC/SnSe composites n – M + MA + HP 0.125 300 ∼7.3 −581 ∼2.5 ∼0.6 −5.77 × 1018 – 2016 [260]
(PbSe)0.9(SnSe)0.1 n – MA + SPS 1.0 773 ∼491.1 ∼−183.3 ∼16.5 ∼1.2 −3.5 × 1019 ∼7.61 2015 [257]
(Sn0.9Se0.87I0.03)(SnS)0.1 n // M + MA + HP ∼1.0 773 ∼16.0 ∼−500.0 ∼4.0 ∼0.36 −3.8 × 1016 (r) 5.75 2016 [259]
Progress in Materials Science 97 (2018) 283–346
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 31. (a) Top view and (b) side view of optimized atomic configuration of monolayer SnSe, reproduced with permission [130], Copyright 2015,
RSC Publishing, (c) two-dimensional charge density of SnSe in the bc plane crossing all Sn atoms, the unit of charge density is e bohr−3, reproduced
with permission [313], Copyright 2016, Nature Publishing Group, and (d) electronic band structures of monolayer SnSe based on PBE (blue dashed
line) and HSE (red solid line) functional and the corresponding DOS [130].

6.1. Monolayer

As discussed above, SnSe is a layered structure [309–311] with a narrow bandgap [197]. A monolayer of SnSe sheet exhibits low
formation energy (about 200 meV per atom) due to the interlayer weak van der Waals force and strong covalent bonds within the
layer [312]. It should be easy to extract a monolayer SnSe sheet from its bulk material without any substrates [130]. In addition,
monolayer SnSe has a large negative Poisson’s ratio of ∼0.17, which may act as a 2D auxetic material [313]. In this respect, SnSe has
full potential to present a high ZT value for a single-layered SnSe sheet [314].

6.1.1. Lattice configuration


Fig. 31(a) illustrates an optimized configuration of monolayer SnSe [130], and shows a highly puckered honeycomb structure
[130,315], which is similar to graphene [316]. Monolayer SnSe possesses Pm21n symmetry with an orthogonal lattice [130], which is
a lower symmetry than that of its bulk counterpart due to the lack of an inversion symmetry [130]. The zigzag (marked as b) and the
armchair (marked as c) axes are denoted in Fig. 31(a). The calculated lattice parameters are 4.307 Å and 4.362 Å respectively along
the b- and c-axes [130]. Fig. 31(b) shows a side view of this configuration. Other results show that the lattice parameters are 4.29 Å
and 4.46 Å respectively along the b- and c-axes [317], which is similar to the results above.
The difference of lattice parameters between b and c may be responsible for the strong anisotropic properties [313], similar to
phosphorene [100,318,319]. The region in Fig. 31(a) characterized by the dashed chart is the primitive unit cell of monolayer SnSe,
which contains two Sn atoms and two Se atoms [130]. Calculations based on DFT show that monolayer SnSe is dynamically stable
[320], and calculations based on AIMD simulations show that monolayer SnSe is thermally stable in a wide temperature range from
300 to 900 K [130].
Fig. 31(c) shows 2D charge density of monolayer SnSe [313] illustrated in the bc plane crossing all Sn atoms. The red and blue
colors depict the high and low charge density, respectively. The unit of charge density is e bohr−3. The electron density exhibits an
obvious symmetry along the b- and c-axes, confirming that the electrons move in square-like potential [313]. The electron density
around Se atoms is obviously higher than that of Sn atoms, indicating that electrons transfer from Sn to Se atoms. It may be owing to
the large difference of electronegativity of Se and Sn atoms, implying that the Sn-Se bond may be a polar covalent bond with a strong

327
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Table 10
A summary of bandgap data of 2D-SnSe and calculation methods.
Composition Structure (Calculated) Bandgap (eV) (Calculation) Method Year Ref.

SnSe Sheet 0.88 First principles Density Functional Theory 2017 [428]
SnSe Monolayer 1.45 mBJ method 2016 [313]
SnSe Monolayer 1.28 Vienna Ab Initio Simulation Package 2015 [130]
SnSe Monolayer 0.898 ∼1.238 Optical-absorption measurements 1990 [429]
SnSe Monolayer 1.79 First principles Density Functional Theory 2017 [384]
SnSe Monolayer 0.99 First principles Density Functional Theory with Vienna Ab Initio Simulation Package 2017 [317]
SnSe Bilayer 0.64, 0.91, 0.82, 0.72 First principles Density Functional Theory 2017 [327]
SnSe Multilayer 0.86 First principles Density Functional Theory 2017 [383]
SnSe 7-layer 1.35 First principles Density Functional Theory 2017 [384]
SnSe 15-layer 1.13 First principles Density Functional Theory 2017 [384]
SnSe Film 1.2 Optical-absorption measurements 2017 [357]
Sn1-xCaxSe Film ∼1.0 (x < 0.5) Optical-absorption measurements 2017 [366]

polarity. Besides, a high electron density is found between the nonbonding neighboring Sn and Se atoms, revealing that strong
nonbonding interactions exist between them, which are weaker than the Sn-Se bond but are still much stronger than common van der
Waals interaction.

6.1.2. Electronic structure


The electronic band structures of monolayer SnSe are different from their bulk counterparts [321]. Fig. 31(d) shows the calculated
the electronic band structure of the monolayer SnSe [130] based on Perdew–Burke–Ernzerhof (PBE) functional within generalized
gradient approximation (GGA) [322] (blue dashed line) and the Heyd–Scuseria–Ernzerhof (HSE06) functional [323] (red solid line),
and the corresponding DOS [130]. The VBM is located on the Γ–Y path, while there exists two energetically nearly degenerate CBMs,
one is on the Γ–Y path, and the other is on the Γ–X path. Both VB and CB have a typical characteristic of pudding mold
[241,324,325], which consists of a dispersive portion with some corrugations and a highly dispersive flat portion, resulting in
multiple Fermi surfaces and in turn contributing to a high S and σ [130].
For the bandgap, calculations based on the GGA–PBE functional suggests that the monolayer SnSe is a quasi-direct-bandgap
semiconductor with a bandgap of 0.914 eV [130], which is larger than that of the bulk SnSe (0.83 eV for α-SnSe and 0.46 eV for β-
SnSe) [179] due to the weak interaction between the layers. Meanwhile, calculations based on the HSE06 functional with 25% Fock
exchange [130] predicted a more accurate bandgap of 1.28 eV for monolayer SnSe, much larger than that calculated using PBE
functional [130]. Such a large bandgap of monolayer SnSe can effectively suppress the bipolar conduction at high temperature, which
may improve its thermoelectric performance [130,326]. By using first-principles DFT calculations, the band structure and bandgap of
a few layer SnSe (such as bilayer SnSe) were also calculated and showed that the interactions between layer play a major role in the
direct-indirect bandgap transition [327]. Table 10 summarizes the bandgap data of 2D SnSe and related calculation methods, which
showing similar values.

6.1.3. Influence of strain


To understand the impact of strain on the band structure of 2D-SnSe, Fig. 32(a) and (b) show the band structures calculated by
mBJ method under uniaxial strain from −10% to 10% along the b- and c-axes [313]. The VBMs (VX and VY) are indicated by small
filled circles and squares, and the CBMs (CX and CY) are indicated by small unfilled circles and squares, respectively. From the
theoretical simulations, with increasing the compressive strain along the c-axis, CX moves down and VY moves up, resulting in a
shrinking indirect bandgap. With increasing the tensile strain along the c-axis, CX moves up and VY moves down, while VX and CY are
nearly stable, resulting in a nearly intact indirect bandgap. With increasing the compressive strain along the b-axis, CX moves up and
VY moves down, while VX and CY are nearly stable, resulting in a nearly intact indirect bandgap. With increasing the tensile strain
along the b-axis, CX moves down and VY moves up, resulting in a shrinking indirect bandgap.
Fig. 32(c) and (d) show the energies of CY, CX, VY, and VX as a function of strain along the b- and c-axes [313]. The compressive
strain along the b-axis affects the same way as the tensile strain along the c-axis, and vice versa. Here, VX is always stable whenever
the compressive and tensile strains are applied along the b- or c-axes. CY acts similarly when the tensile strain along the c-axis or the
compressive strain along the b-axis is applied. VY and CX are always sensitive and have nearly linear response to any strain, showing
an anisotropic response of electronic structures to the strain [313]. In Fig. 32(c) and (d), the colored areas marked by II and V
correspond to the direct bandgap, while I, III, IV and VI represent the indirect bandgap. A strain-induced direct - indirect bandgap
transition can be observed. It is clear that a low strain can give rise to diverse electronic structures, which makes 2D-SnSe an excellent
2D semiconductor material for the strain band engineering [328,329].

6.1.4. Performance forecasting


Similar to bulk (3D) SnSe thermoelectric materials, modelling is essential for predicting thermoelectric performance for 2D-SnSe
sheets. Fig. 33(a)–(c) show the calculated S, σ, and S2σ as a function of chemical potential (μe) along different directions at different
temperatures [130], in which the calculations were based on BTE using the BoltzTraP code [130]. Fig. 33(a) shows that S varies
dramatically with μe, indicating that an optimal n is beneficial to achieve efficient thermoelectric performance. S decreases rapidly

328
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 32. Calculated band structures by mBJ method under uniaxial strain from -10% to 10% along (a) c- and (b) b-axis, the CBM and VBM are
marked by CY, CX and VY, VX, respectively. The energies of CY, CX, VY, and VX as a function of strain in the (c) c- and (d) b-axis are also plotted.
Reproduced with permission [313]. Copyright 2016, Nature Publishing Group.

with increasing the temperature due to a bipolar conduction effect or the excitation of carriers with both positive and negative
charges, suggesting an increasing inverse Hall coefficient [130]. The peak S for n-type monolayer SnSe were - 2150 μV K−1 along the
b-axis and - 2080 μV K−1 along the c-axis at room temperature [130], and the peak S for p-type monolayer SnSe were 2160 μV K−1
along the b-axis and 2200 μV K−1 along the c-axis at room temperature [130], respectively. These enhanced S nearly four times when
compared with that of the bulk SnSe [61]. Such remarkable enhancement of S can be ascribed to the quantum confinement effect
which induces sharp DOS peaks near the Fermi energy [130]. Fig. 33(b) shows that σ is less sensitive to temperature [130]. With
increasing μe, σ increases drastically. Simulations based on extensive ab initio calculations show a very high hole mobility (10000 cm2
V−1 s−1) for monolayer SnSe [313]. Consequently, as shown in Figure 33(c), S2σ of monolayer SnSe (∼250 μW cm−1 K−2 along the
b-axis and ∼275 μW cm−1 K−2 along the c-axis at 700 K) [130] were much larger than that of bulk SnSe materials [61]. Other
calculations based on DFT combined with Boltzmann transport theory showed that for SnSe monolayer, along the c-axis, the carrier
mobility μx was 1035.84 cm2 V−1 s−1, 621.5 cm2 V−1 s−1, and 443.93 cm2 V−1 s−1 at 300 K, 500 K, and 700 K, respectively, and the
effective mass m∗x was 0.15 me [317]; along the b-axis, the carrier mobility μy was 965.23 cm2 V−1 s−1, 579.13 cm2 V−1 s−1, and
413.6 cm2 V−1 s−1 at 300 K, 500 K, and 700 K, respectively, and the effective mass m∗y was 0.16 me, which was close to that of along
the c-axis [317]. The calculated peak S was 1750 μV K−1 at 300 K [317].
For the calculations of thermal transport properties of monolayer SnSe, Fig. 33(d) shows that κe exhibits trends similar to σ due to
the relationship of κe = LσT, where L was approximated to 1.5 × 10−8 W Ω K−2 in the case of non-degenerate semiconductors
limited by phonon scattering [241]. Fig. 33(e) shows the temperature dependence of axial κl, which is satisfied with the relationship
of κl ∝ T−1, indicating that anharmonic phonon–phonon interactions are dominant in the phonon scattering [330]. The calculated κl
was ∼2.5 W m−1 K−1 along the b-axis [130] and ∼2.02 W m−1 K−1 along the c-axis at room temperature [130]. Extensive ab initio
calculations showed that SnSe monolayer possessed an ultralow κl (<3 W m−1 K−1 at 300 K), making it probably the most heat-
insulating and flexible material in known 2D atomic materials [313]. The low intrinsic κ is a combined consequence of the heavy
constituent elements and the low atomic coordination [130]. The calculated average γ values along the b- and c-axis were 1.65 and
2.03, respectively [130]. However, it should be noticed that the κl calculated for a monolayer SnSe [130] was much higher than that
of the bulk counterpart [61,65]. This is because the real bulk SnSe has many other scattering sources, such as boundary, defect, and
impurity, while monolayer SnSe loses the distorted SnSe7 polyhedra structure [61], weakening the anharmonicity. Fig. 33(f) shows
the ZT value of monolayer SnSe as a function of μe at different temperatures [130]. At 700 K, the peak ZT values of 2.47 at μe =
−0.62 eV and 2.76 at μe = 0.575 eV along the b-axis [130], and the peak ZT values of 2.51 at μe = −0.63 eV and 3.27 at μe = 0.54 eV
along the c-axis were predicted [130]. The calculated ultrahigh peak ZT values indicated a great potential for monolayer SnSe as a 2D
high performance thermoelectric material [130].

329
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

16
a 2
700 K, b b 700 K, b
500 K, b 500 K, b
300 K, b 300 K, b
12

S (103 μV K-1)
700 K, c

(106 S m-1)
1 700 K, c
500 K, c 500 K, c
300 K, c 300 K, c
0 8

-1
4

-2
0
-2 -1 0 1 2 -2 -1 0 1 2
e
(eV) e
(eV)
c 30 700 K, b d 75
700 K, b
500 K, b 500 K, b
25 300 K, b 300 K, b
S2 (mW m-1K2)

700 K, c 60 700 K, c

(W m-1K-1)
20 500 K, c 500 K, c
300 K, c 45 300 K, c
15
30
10 e

5 15

0 0
-2 -1 0 1 2 -2 -1 0 1 2
e
(eV) e
(eV)

e 2.7
b f 700 K, b 500 K, b
4 300 K, b 700 K, c
2.4 c 500 K, c 300 K, c
2.1 3
(W m-1K-1)

3.27

1.8
ZT

2
1.5
l

1.2 1
0.9
0
300 400 500 600 700 -2 -1 0 1 2
T (K) e
(eV)
Fig. 33. Calculated chemical potential-dependent (a) S, (b) σ, (c) S2σ and (d) κe at different temperature, (e) calculated temperature-dependent κl
and (f) calculated chemical potential-dependent ZT at different temperature. Reproduced with permission [130]. Copyright 2015, RSC Publishing.

6.2. Recent advances

Until now, it has been successfully fabricated macro-sized 2D-SnSe materials (such as SnSe film), however its application is still a
challenge in the thermoelectric device. For micro/nano-sized 2D-SnSe materials (such as SnSe nanosheets [331–334]), both the
fabrication technique and the application are under exploration. This section will discuss the recent advances in the fabrications and
applications of 2D-SnSe materials.

6.2.1. Films
For fabricating SnSe films, the synthetic methods are relatively mature, including atomic layer deposition (ALD) [335], brush
plating technique [336], simple bath deposition [337–339], thermal evaporation method [340–346], solution method [347,348], hot
wall epitaxy technique [349–351], flash evaporation method [352,353], vacuum deposition technique [309,354–357], chemical
deposition technique [358], pulsed laser deposition (PLD) [359–366], spray pyrolysis technique [367], and electrodeposition
technique [368–372]. Most of the fabricated films are orthorhombic α-SnSe while rock-salt structure (1 1 1) SnSe film was also grown
with molecular beam epitaxy (MBE) technique on Bi2Se3 surface [273]. Fig. 34(a) reveals XRD patterns of a bare SrTiO3 (STO)

330
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Fig. 34. (a) XRD patterns of a bare STO substrate, a 5 QL Bi2Se3 film on STO substrate and a 16 nm SnSe film grown on Bi2Se3 film on STO, (b) STM
image of a 7 nm SnSe film (60 nm × 60 nm), the inset shows the line profile along the black line, (c) Cross-section HRTEM image of a 16 nm SnSe
film grown on 5 QL Bi2Se3 film. Reproduced with permission [273]. Copyright 2015, John Wiley and Sons.

substrate, a 5 quintuple-layer (QL) Bi2Se3 film on STO substrate and a 16 nm SnSe film grown on Bi2Se3 film on STO [273]. Fig. 34(b)
shows the STM image of a 7 nm SnSe film (60 nm × 60 nm) [273]. The inset shows the line profile along the black line. The STM-
measured step height, well consistent with the peak, corresponds to a spacing d = 3.37 Å shown in Fig. 34(a). Fig. 34(c) shows the
cross-section HRTEM image of a 16 nm SnSe film grown on 5 QL Bi2Se3 film [273]. The SnSe film shows an obvious bilayer structure
with the lattice constants along the [1 1 1] and [1 1 0] directions being 4.19 Å and 3.37 Å, respectively [273].
Fig. 35(a) shows the XRD patterns taken from SnSe thin films deposited at a normal angle and at an 80° glancing angle [359]. For
the thin film grown at a normal angle, the strong diffraction peaks of (2 0 0), (4 0 0) and (8 0 0) indicate that SnSe films were
preferentially grown along the a-axis during the depositing process. For the thin film grown with a glancing angle, (2 0 0) and (8 0 0)
diffraction peaks are not present, suggesting that the growth direction and orientation is not preferred. Fig. 35(b) and 35(c) shows the
cross sectional SEM images of film deposited at normal angle and glancing angles [359], respectively. It is clear that the film
deposited at a normal angle possesses large grains, while the film grown at a glancing angle shows a nanopillar structure rather than a
continuous film on the substrate, and the nanopillars are grown with a relatively high density. The thickness of SnSe thin film grown

d 300
Normal angle
80° glacing angle

200
σ (S cm-1)

100

300 350 400 450 500 550


T (K)

e 600 Normal angle f 24 Normal angle g 0.20 Normal angle


80° glacing angle 80° glacing angle 80° glacing angle
S2σ (μW cm-1K-2)

500 20
0.18
κ (W m-1K-1)

400 16
S (μV/K)

300 0.16
12
200 8 0.14
100 4 0.12
0 0
0.10
350 400 450 500 550 300 350 400 450 500 550 200 225 250 275 300 325
T (K) T (K) T (K)
Fig. 35. (a) XRD patterns of the thin film deposited at a normal angle and an 80° glancing angle with Ts = 603 K, cross sectional SEM images of films
deposited at (b) normal angle and (c) glancing angle at Ts = 603 K, and temperature-dependent thermoelectric properties for thin SnSe film
deposited at different angles: (d) σ, (e) S, (f) S2σ, and (g) κ. Reproduced with permission [359]. Copyright 2017, Elsevier Publishing.

331
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

a 108 cm-1 Raman spectrum b 0.15 c 8000 S


500
-1
Intensity (a.u.)

130 cm
0.14 6000
400

(S cm-1)

S ( V K-1)
150 cm-1
0.13 4000
300
0.12 2000
200
0.11
0
50 100 150 200 250 0 50 100 150 200 250 300 0 50 100 150 200 250 300
-1
Raman shift (cm ) T (K) T (K)
d 7 S2 e 4.5 f ZT
1.2
S2 (10-4 W m-1K-2)

6
(10-2 W m-1K-1)
4.0
5
4 0.8
3.5

ZT
3
3.0 0.4
2
1 2.5
0.0
0
2.0
0 50 100 150 200 250 300 0 50 100 150 200 250 300 0 50 100 150 200 250 300
T (K) T (K) T (K)
Fig. 36. (a) Raman spectrum of the SnSe thin film [357], and temperature-dependent thermoelectric properties for thin SnSe film: (b) σ, (c) S, (d)
S2σ, (e) κ and (f) ZT. Reproduced with permission [357]. Copyright 2017, IOP Publishing.

with the normal angle can be estimated as ∼730 nm, while it is ∼300 nm for its counterpart [359].
In terms of the property measurements (such as S, σ, and κ), the Harman transient method [373–375] and the time-domain
thermos-reflectance (TDTR) method [376–378] were widely used. To study the thermoelectric performance of SnSe films deposited at
different angles, Fig. 35(d)–(g) show σ, S, S2σ and out-of-plane κ [359]. It suggested that the film deposited at a glancing angle
possesses much better σ, S, and S2σ than that deposited at normal angle with the entire temperature, as well as lower κ before 300 K. A
high S of 498.5 μV K−1 and a high S2σ of 18.5 μW cm−1 K−2 [359] were achieved. The results reveal that the glancing angle
deposition can greatly reduce the grain size of the thin film and significantly enhance S and S2σ. The enhanced thermoelectric
property can be attributed to the potential barrier scattering at grain boundaries owing to the reduced grain size and increased grain
boundaries in the film [359].
Other studies showed that p-type SnSe thin films possess good thermoelectric performance at low temperature [357]. Fig. 36(a)
shows Raman spectrum of the SnSe thin film [357]. The peak at 108 cm−1 corresponds to B3g vibration mode and the peaks at 130
and 150 cm−1 both correspond to Ag vibration mode, which all belong to α-SnSe [379]. Fig. 36(b)–(f) show temperature-dependent
thermoelectric properties for thin SnSe film [357]. At 42 K, SnSe thin film exhibited an ultrahigh S of ∼7863 μV K−1, as well as a
good S2σ of ∼7.2 μW cm−1 K−2, resulting in a high ZT of ∼1.2 [357]. The ultrahigh S observed below 84 K should derive from the
effect of phonon drag on charge carriers, which results from the interaction of phonons with mobile carriers and both the electrons
and phonons contribute to the remarkable rise in S [357,380]. These results showed that SnSe thin film can be considered as an ideal
material for low temperature thermoelectric applications. For n-type SnSe thin film [381,382], a new cross-plane thermoelectric
measurement was conducted and applied on the measurement of (SnSe)x(TiSe2)x (x = 1, 3, 4, 5) [381], the cross-plane S decreased
from −31 to −2.5 μV K−1 with x decreased from 5 to 1 [381]. SnSe films prepared by spray pyrolysis technique [367] also showed a
low thermoelectric performance. More studies are undertaking to further improve the properties of n-type SnSe thin films.

6.2.2. Nanosheets
Nanosheets generally possess thicknesses of multiple Sn-Se monolayers and each monolayer has a thickness of 5.75 Å [383]. For
SnSe nanosheets with fewer layers (such as bilayer [384]), historically, the phase-controlled synthesis is a challenge [348]. In 2013, a
single-crystalline SnSe nanosheet with four-atomic thickness was synthesized for the first time [385]. In 2014, a phase-controlled
method was employed to synthesize SnSe nanosheets [348]. Since then, significant progress has been made in fabricating SnSe
nanosheets and exploring their thermoelectric applications.
By a two-step synthesis method, synthesized first by Chemical vapor deposition (CVD) [386–388] method and subsequently
followed by a nitrogen etching technique, both multi-layered nanosheets and monolayer SnSe were obtained on the SiO2/Si substrate
[334]. Fig. 37(a) shows the AFM image and inserted SEM image of rectangular shaped SnSe nanosheets with lateral dimensions of
about 30 μm × 50 μm, as well as a thickness of 54.9 nm, and Fig. 37(b) shows the corresponding images of a SnSe monolayer with a

332
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

d 110 SnSe/Ag contact SnSe/Au contact e 150 -50 V f 20 -80 V


15
100 100 50 V
10 80 V
90 50
Id (nA)

Id (nA)

Id (nA)
80 0 0
-50 -5
70
-10
60 -100
SnSe/Ag contact -15
50 SnSe/Au contact -150 -20
-80 -40 0 40 80 -2 -1 0 1 2 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
Vg (V) Vds (V) Vds (V)
Fig. 37. AFM image and inserted SEM image of (a) SnSe nanosheet with a thickness of 54.9 nm and (b) monolayer with a thickness of 6.8 Å, (c)
HRTEM image of the monolayer with inserted SAED pattern, (d) Transport characteristic curves of the devices at Vds = 1 V for the SnSe/Ag and
SnSe/Au contacts, respectively, insets show the corresponding optical images of the devices, and output characteristic curves of the devices with the
(e) SnSe/Ag and (f) SnSe/Au contacts at different Vg values, respectively. Reproduced with permission [334]. Copyright 2017, IOP Publishing.

thickness of 6.8 Å (tested by AFM), which is close to the theoretical value [334]. Fig. 37(c) shows a HR-TEM image of monolayer SnSe
with inserted SAED pattern [334], indicating that the obtained monolayer has a typical orthorhombic structure of α-SnSe.
In terms of the electrical transport properties, electrical devices using individual rectangular shaped SnSe nanosheet were fab-
ricated on a 300 nm SiO2/Si substrate, and the 50 nm Au and 50/10 nm Ag/Au contact electrodes were used [334]. Fig. 37(d) show
transport characteristic curves of the devices at Vds = 1 V for the SnSe/Ag and SnSe/Au contacts, respectively; and insets show the
corresponding optical images of the devices [334]. Both of them exhibited p-doped ambipolar characteristics. The hole field-effect
mobility at room temperature was estimated as ∼0.11 and 0.16 cm2 V−1 s−1 for Ag-contacted and Au-contacted transistors, re-
spectively; which are comparable to some other 2D-thermoelectric materials such as ultrathin SnSe2 nanosheets (∼0.6 cm2 V−1 s−1)
[389] and multilayer GaTe nanosheets (∼0.1 cm2 V−1 s−1) [390]. Fig. 37(e) and (f) show the output characteristic curves of the
devices made by the SnSe/Ag and SnSe/Au contacts at different Vg values, respectively [334]. The linear increase of the drain current
(Id) with drain voltage (Vds) suggests an excellent ohmic contact at the source and drain electrode. These linear output curves also
indicate the high electronic quality of the samples, and the device resistance of the nanosheets is estimated as ∼15 MΩ [334].
Except for the electrical transport properties, the study of temperature-dependent Raman spectroscopy behavior shows that SnSe
nanosheets have a high susceptibility to temperature [386], as well as obvious anharmonic phonon properties [386]. Calculations
based on first-principle DFT showed that the bandgap monotonically increased with the reduction of the nanosheet thickness [384],
which indicated full potentials for use in thermoelectric applications. For other cases of the thermoelectric performance, SnSe na-
nosheets were obtained from a bulk SnSe ingot by pulverization, Li-intercalation, and exfoliation processes [391–394] in 2016, and a
peak ZT of 0.32 was obtained from this SnSe nanosheets [391]. Similar, SnSe1-xSx nanosheets are prepared from bulk by hydro-
thermal Li intercalation and subsequent exfoliation [331,392–395]. The scattering of phonons at a number of atomic disorders and
nanosized boundaries derived from the substitution of S atoms into SnSe and contributed to an effective reduction of κ and an
improved peak ZT of 0.12 at 310 K [331]. The thermoelectric performance of 2D nanostructured SnSe still needs further improvement
in their synthesis, characterization and thermoelectric performance.

6.2.3. Anisotropy behavior


In terms of the polarized Raman response and electrical transport studies, Fig. 38(a) shows a measured optical image of a
measured nanosheet, in which the crystallographic axes (y represents b-axis and z represents c-axis), the incident light polarization
direction, and the angle θ between the incident light are indicated [387]. Fig. 38(b) and (c) show the typical polarized Raman spectra
of a SnSe nanosheet with different sample rotation angles excited using a 633 nm laser under parallel-polarization configuration and
cross-polarization configuration [387], respectively. The polarized Raman spectra exhibit an angular dependence that reveals the

333
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

b Ag1
B3g
Ag2 c Ag1 B3g
Ag2
Ag3 180o
180o

Intensity (a.u.)
Intensity (a.u.)
150 o 150o

120 o 120o
90 o
90o
60 o
60o
30 o
30o
o
0 0o

40 80 120 160 200 40 80 120 160 200


Raman shift (cm-1) Raman shift (cm-1)

Fig. 38. (a) Optical image of a measured SnSe nanosheet. Typical polarized Raman spectra of a nanosheet with different sample rotation angles
excited using a 633 nm laser under (b) parallel-polarization configuration and (c) cross-polarization configuration, (d) side views of the atomic
displacements of the four modes measured in the 50–200 cm−1 range, the two rows in the figure represent two adjacent layers. The calculated
frequencies provided in parentheses are in good agreement with the experimental values. (e) Optical image of a typical nanosheet with 12 electrodes
spaced at 30° apart, and (f) the polar plot of the angle-resolved conductivity. Reproduced with permission [387]. Copyright 2016, RSC Publishing.

anomalous anisotropic light-matter interaction of the crystal [387]. Because only four Raman modes (Ag1, B3g, Ag2, and Ag3 modes)
can be resolved in this case [387], Fig. 38(d) shows the side views of the atomic displacements of these four modes, in which the two
rows in the figure represent two adjacent layers [387]. The calculated frequencies provided in parentheses are in good agreement
with the experimental values [387]. Fig. 38(e) shows the optical image of a typical SnSe nanosheet with 12 electrodes spaced at 30°
apart [387], and Fig. 38(f) shows the polar plot of the angle-resolved conductivity [387]. The angle-dependent conductivity shows
strong anisotropy, with a maximum conductivity along the zigzag direction and a secondary maximum conductivity along the
armchair direction.
Other studies also demonstrated that SnSe nanosheets have obvious anisotropic electrical transport properties [383,388]. A large
anisotropic ratio of carrier mobility μx /μy of ∼5.8 between the b- and c-axes was obtained, which was a record high value reported
for 2D anisotropic materials [383].

6.2.4. Organic/inorganic composites


Organic polymer thermoelectric materials, such as polypyrrole [396,397], polythiophene [391,398], and polyaniline [399,400],
are promising candidates due to their relatively high σ and intrinsically low κ. Among them, poly (3, 4-ethylenedioxythiophene):poly
(styrenesulfonate) (PEDOT:PSS) is the most promising polythiophene-based material, which possesses remarkable high σ [401–404].
However, when compared with inorganic materials, the performance of PEDOT:PSS is restricted by relatively low S2σ deriving from
low S [405]. To overcome these issues, inorganic fillers with high S, such as Bi2Te3 [406,407] and PbTe [408], has been added into
the system to improve S of conducting polymer-based thermoelectric materials, resulting in a kind of inorganic/organic composites.
Among these inorganic candidates, SnSe possesses relative high S, high σ and very low κ, which is a good candidate for the role of a
composite material.
To illustrate, Fig. 39(a) shows a schematic representation of the overall fabrication procedure for various SnSe/PEDOT:PSS
composites [332]. SnSe nanosheets were obtained from the SnSe ingot after Li-intercalation and following exfoliation [392,393,409],
and then evenly distributed in the PEDOT:PSS matrix and affected the intrinsic conduction of the composites. In addition, the polar
solvent treatment [410–412] applied to the prepared SnSe/PEDOT:PSS composites can further improve the thermoelectric perfor-
mance [332]. Fig. 39(b)–(d) show the thermoelectric properties (σ, S, S2σ and κ) as a function of the percentage of SnSe nanosheets
for SnSe/PEDOT:PSS composite films [332]. With increasing the SnSe nanosheet content, although σ decreases to some extent, the S
of SnSe/PEDOT:PSS composites increase significantly, leading to a high S2σ at high SnSe content. σ and S can also be modeled as
series-and parallel-connected composites in the simple case of a filler/matrix binary system [413,414]. Meanwhile, κ also increased
from ∼0.22 W m−1 K−1 to 0.6 W m−1 K−1 [332]. To further enhance the thermoelectric performance of SnSe/PEDOT:PSS

334
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

b 20 SnSe/PEDOT:PSS c 600 SnSe/PEDOT:PSS d κ


Parallel-connected model Parallel-connected model
18 200 S2σ 0.6
Series-connected model 500 Series-connected model S2σ (μW m-1K-2)

κ (W m-1K-1)
16
σ (S cm-1)

S (μV K-1)

400
14
300 100 0.4
12
200
10
8 100
0 0.2
6 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Weight Fraction of SnSe Weight Fraction of SnSe Weight Fraction of SnSe

e 1000 SnSe_0 SnSe_10 f g 500 SnSe_0 SnSe_10


SnSe_20 SnSe_30 600 SnSe_20 SnSe_30
800 SnSe_50 SnSe_50 SnSe_100
400
S2σ (μW m-1K-2)
S (μV K-1)

SnSe_100
σ (S cm-1)

SnSe_0 SnSe_10
600 400 SnSe_20 SnSe_30 300
SnSe_50 SnSe_100
400 200
200
200 100

0 0
0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Solvent content (vol.%) Solvent content (vol.%) Solvent content (vol.%)
Fig. 39. (a) Schematic diagram of the overall fabrication procedure for SnSe/PEDOT:PSS composite film, and thermoelectric properties as a function
of the SnSe nanosheet content for SnSe/PEDOT:PSS composite films: (b) σ, (c) S, (d) S2σ and κ, and thermoelectric properties as a function of the
solvent (DMSO) content for varied SnSe/PEDOT:PSS composites: (e) σ, (f) S, and (g) S2σ. Reproduced with permission [332]. Copyright 2016,
Elsevier Publishing.

composites, various proportions of organic solvent (DMSO) were mixed with the different amounts of SnSe nanosheets dispersed
PEDOT:PSS solutions. Fig. 39(e)–(g) show thermoelectric properties (σ, S, S2σ and κ) as a function of the DMSO content for varied
SnSe/PEDOT:PSS composites [332]. After the DMSO treatment, the SnSe/PEDOT:PSS composites exhibited a significant σ en-
hancement without considerable reduction of the S values. In addition, the composites with large amounts of PEDOT:PSS exhibited
enhanced σ values, which mainly derived from the morphological changes of PEDOT:PSS in the composite structure occurring upon

335
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

addition of the polar solvents [415]. As a result, S2σ of DMSO-treated composite films were considerably higher than that of the
composite film without solvent treatment [332]. The 5 vol% DMSO-treated SnSe_20/PEDOT:PSS composite film had the maximum
S2σ of ∼3.86 μW cm−1 K−2, which is ∼257 times higher than that of the pristine PEDOT:PSS film (∼0.015 μW cm−1 K−2) [332].

7. Conclusion, challenge, and strategy

Compared with other traditional thermoelectric materials, SnSe-based thermoelectric materials are highly promising. SnSe
consists of light elements and has a very small and simple unit cell, which is against the established rules for searching a promising
thermoelectric material [64]. The record high ZT of ∼2.6 at 923 K for single crystals along the b-axis [61] was achieved without
optimizing the n, resulting in a moderate S2σ of 8.5 μW cm−1 K−2 at 923 K [61]. An ultralow κ of only ∼0.35 W m−1 K−1 at 923 K is
attributed to the strong anharmonicity of chemical bonds in this SnSe layered compound [61]. Thus, the finding of SnSe was a real
breakthrough in the thermoelectric community and has a guidance to exploring and designing high-efficiency thermoelectric ma-
terials with low κ.
In terms of SnSe single crystals, suffered from the poor mechanical properties, rigid crystal growth conditions, and high cost of
their production, single crystals are difficult to be synthesized with a large scale and a high success rate [65], it is still a restriction for
the development of high-performance single crystals and their practical application. Besides, further investigations are necessary to
solve the controversies on whether the ultralow κ values reported of SnSe single crystals [61] are intrinsic and their high-temperature
thermoelectric properties after phase transition. Furthermore, when doping is applied in single crystal SnSe, the inappropriate type
and/or amount of dopant may cause a failure of the growth of big single crystals, which is still a big problem for the accurate
thermoelectric evaluations. As for polycrystalline SnSe, it is hard to achieve a very strong anisotropy like single crystal because the
SnSe powders (grains) are hard to array orderly along the uni-axis by sintering, leading to a much lower S2σ than that of the single
crystals. The effect of doping on polycrystalline SnSe is limited because it is still a challenge to optimizing n to achieve a maximized
ZT as indicated by theoretical calculation. Last but not least, the mechanism of inducing inclusions to further improve the ther-
moelectric performance of polycrystalline SnSe is far to fully understand.
In this respect, we point out potential strategies to dealing with the challenges for further improving the thermoelectric per-
formance of SnSe-based thermoelectric materials and summarize in the compass-like diagram of Fig. 40. Because SnSe single crystals
possess the highest ZT along b-axis, optimization of fabrication process is highly needed for the growth of single crystal with re-
peatable and trustable quality. In particular, it is urgently to find new facile methodology to fabricate high-quality SnSe single crystal
for the assembly of thermoelectric devices and their evaluations, which will expand the real application of SnSe. For polycrystalline
SnSe, there are mainly three routes to further increase its thermoelectric performance. For the texturing, it is better to synthesize the
polycrystalline SnSe with both large scale and high crystallinity along b- or c-axis for realization of high anisotropy. Facile solve-
thermal method has been demonstrated to possess the ability to large-scale fabrication of SnSe plates with their lateral dimension of
up to hundreds µm along b- or c-axis and controllable thickness from several nm to tens of nm [91]. After advanced sintering process,
a very strong anisotropy can be achieved in the products fabricated by solve-thermal method, leading to a good thermoelectric
performance. Therefore, it is high potential to expanding the fabrication research on solve-thermal method. Using nanostructuring to
tune grain size of polycrystalline SnSe will help to control the density of grain boundary and crystal defects, contributing to
strengthening the phonon scattering and further reducing κ. Both anion and cation doping or alloying are suitable to optimizing n and
band structures and in turn enhance their thermoelectric properties.

Fig. 40. Compass-like illustration of the ZT enhancement strategies for SnSe-based thermoelectric materials.

336
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

In particular, four strategies will be highlighted here for the future development of SnSe. The first strategy is the combination of
band engineering and texturing to maximize the anisotropy and S2σ with optimized n [277]. Second strategy is the combination of
band engineering and single crystal growth. Both p-type [78] and n-type [52] doping have showed the potentials to improving the
thermoelectric performance of single crystals. This is an important way to further improve the thermoelectric performance of α-SnSe,
as well as realizing n-type doping with a high thermoelectric performance. Third strategy is the combination of nanostructuring and
texturing. Nanograins possess the same axis (along b- and c-axis) via ideal texturing process, κ along these axes will have ultralow
values, which may be even lower than that of single crystals due to the addition of phonon scattering effects occurring at grain
boundaries and crystal defects. In this respect, a very high ZT can be predicted. Fourth strategy is the combination of nanostructuring
and single crystal growth. Crystal defects such as point defects can be induced into the single crystals to further reduce the κ by
strengthening the phonon scattering effect. Thus, an ideal SnSe-based thermoelectric material should be described as polycrystalline
SnSe composed by nanograins with single-crystal-like anisotropy and optimized n.

Acknowledgement

The authors are thankful to Lei Yang, Min Hong, Yichao Zou, Lina Cheng, Guang Han, Weidi Liu, Yuan Wang, Raza Moshwan,
Shengduo Xu and Lijun Wang for their respective valuable experimental contributions and constructive discussion. This work was
financially supported by the Australian Research Council. ZGC thanks the USQ Start-up Grant and Strategic Research Projects. LDZ
thanks the Beijing Municipal Science & Technology Commission under Grant No. Z17110002017002. The Australian Microscopy &
Microanalysis Research Facility is acknowledged for providing characterization facilities.

References

[1] Bell LE. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008;321:1457–61.
[2] Sootsman JR, Chung DY, Kanatzidis MG. New and old concepts in thermoelectric materials. Angew Chem Int Ed 2009;48:8616–39.
[3] Snyder GJ, Toberer ES. Complex thermoelectric materials. Nat Mater 2008;7:105–14.
[4] Poudel B, Hao Q, Ma Y, Lan Y, Minnich A, Yu B, et al. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science
2008;320:634–8.
[5] Dresselhaus MS, Chen G, Tang MY, Yang RG, Lee H, Wang DZ, et al. New directions for low-dimensional thermoelectric materials. Adv Mater 2007;19:1043–53.
[6] Biswas K, He J, Blum ID, Wu CI, Hogan TP, Seidman DN, et al. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature
2012;489:414–8.
[7] Liang J, Li F, Cheng H-M. On Energy: clean conversion and smart storage in the future. Energy Storage Mater 2016;3:A1–2.
[8] Chen Z-G, Han G, Yang L, Cheng L, Zou J. Nanostructured thermoelectric materials: current research and future challenge. Prog Nat Sci 2012;22:535–49.
[9] Han G, Chen ZG, Drennan J, Zou J. Indium selenides: structural characteristics, synthesis and their thermoelectric performances. Small 2014;10:2747–65.
[10] Qiu P, Shi X, Chen L. Cu-based thermoelectric materials. Energy Storage Mater 2016;3:85–97.
[11] Moshwan R, Yang L, Zou J, Chen Z-G. Eco-friendly SnTe thermoelectric materials: progress and future challenge. Adv Funct Mater 2017;27:1703278.
[12] Gayner C, Kar KK. Recent advances in thermoelectric materials. Prog Mater Sci 2016;83:330–82.
[13] Heremans JP, Jovovic V, Toberer ES, Saramat A, Kurosaki K, Charoenphakdee A, et al. Enhancement of thermoelectric efficiency in PbTe by distortion of the
electronic density of states. Science 2008;321:554–7.
[14] Zhang Q, Wang H, Liu W, Wang H, Yu B, Zhang Q, et al. Enhancement of thermoelectric figure-of-merit by resonant states of aluminium doping in lead selenide.
Energy Environ Sci 2012;5:5246–51.
[15] Hong M, Chen ZG, Yang L, Zou YC, Dargusch MS, Wang H, et al. Realizing zT of 2.3 in Ge1−x−ySbxInyTe via Reducing the Phase-Transition Temperature and
Introducing Resonant Energy Doping. Adv Mater 2018. 1705942.
[16] Hong M, Chen ZG, Yang L, Liao ZM, Zou YC, Chen YH, et al. Achieving zT > 2 in p-Type AgSbTe2−x sex alloys via exploring the extra light valence band and
introducing dense stacking faults. Adv Energy Mater 2017. 1702333.
[17] Yang H, Bahk J-H, Day T, Mohammed AM, Snyder GJ, Shakouri A, et al. Enhanced thermoelectric properties in bulk nanowire heterostructure-based nano-
composites through minority carrier blocking. Nano Lett 2015;15:1349–55.
[18] Burke P, Bowers J. Incorporating minority charge carrier blocking regions with thermoelectric materials for improved thermoelectric device efficiency, vol. 20.
University of California Invention Disclosure; 2013.
[19] Bahk J-H, Shakouri A. Minority carrier blocking to enhance the thermoelectric figure of merit in narrow-band-gap semiconductors. Phys Rev B 2016;93:165209.
[20] Burke PG, Curtin BM, Bowers JE, Gossard AC. Minority carrier barrier heterojunctions for improved thermoelectric efficiency. Nano Energy 2015;12:735–41.
[21] Pei Y, Shi X, LaLonde A, Wang H, Chen L, Snyder GJ. Convergence of electronic bands for high performance bulk thermoelectrics. Nature 2011;473:66–9.
[22] Liu W, Tan X, Yin K, Liu H, Tang X, Shi J, et al. Convergence of conduction bands as a means of enhancing thermoelectric performance of n-type Mg2Si(1-x)Sn(x)
solid solutions. Phys Rev Lett 2012;108:166601.
[23] Liu X, Zhu T, Wang H, Hu L, Xie H, Jiang G, et al. Low electron scattering potentials in high performance Mg2Si0.45Sn0.55 based thermoelectric solid solutions
with band convergence. Adv Energy Mater 2013;3:1238–44.
[24] Lv H, Lu W, Shao D, Sun Y. Enhanced thermoelectric performance of phosphorene by strain-induced band convergence. Phys Rev B 2014;90:085433.
[25] Dresselhaus M, Chen G, Tang M, Yang R, Lee H, Wang D, et al. New directions for nanoscale thermoelectric materials research. MRS Proceedings 2011;886.
[26] Tian Y, Sakr MR, Kinder JM, Liang D, MacDonald MJ, Qiu RL, et al. One-dimensional quantum confinement effect modulated thermoelectric properties in InAs
nanowires. Nano Lett 2012;12:6492–7.
[27] Harman T, Taylor P, Walsh M, LaForge B. Quantum dot superlattice thermoelectric materials and devices. Science 2002;297:2229–32.
[28] Harman T, Spears D, Manfra M. High thermoelectric figures of merit in PbTe quantum wells. J Electron Mater 1996;25:1121–7.
[29] Hong M, Chasapis TC, Chen Z-G, Yang L, Kanatzidis MG, Snyder GJ, et al. N-Type Bi2Te3–xSex nanoplates with enhanced thermoelectric efficiency driven by
wide-frequency phonon scatterings and synergistic carrier scatterings. ACS Nano 2016;10:4719–27.
[30] Yang L, Chen Z-G, Hong M, Wang L, Kong D, Huang L, et al. N-type Bi-doped PbTe nanocubes with enhanced thermoelectric performance. Nano Energy
2017;31:105–12.
[31] Yang L, Chen ZG, Hong M, Han G, Zou J. Enhanced thermoelectric performance of nanostructured Bi2Te3 through significant phonon scattering. ACS Appl
Mater Interf 2015;7:23694–9.
[32] Yang L, Chen Z-G, Han G, Hong M, Zou Y, Zou J. High-performance thermoelectric Cu2Se nanoplates through nanostructure engineering. Nano Energy
2015;16:367–74.
[33] Han G, Chen Z-G, Yang L, Hong M, Drennan J, Zou J. Rational design of Bi2Te3 polycrystalline whiskers for thermoelectric applications. ACS Appl Mater Interf
2014;7:989–95.
[34] Jood P, Ohta M. Hierarchical architecturing for layered thermoelectric sulfides and chalcogenides. Materials 2015;8:1124–49.

337
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[35] Bhattacharya S, Bohra A, Basu R, Bhatt R, Ahmad S, Meshram K, et al. High thermoelectric performance of (AgCrSe2)0.5(CuCrSe2)0.5 nano-composites having
all-scale natural hierarchical architectures. J Mater Chem A 2014;2:17122–9.
[36] Liu W-S, Zhang Q, Lan Y, Chen S, Yan X, Zhang Q, et al. Thermoelectric property studies on Cu-doped n-type CuxBi2Te2.7Se0.3 nanocomposites. Adv Energy
Mater 2011;1:577–87.
[37] Pei Y, Lensch-Falk J, Toberer ES, Medlin DL, Snyder GJ. High thermoelectric performance in PbTe due to large nanoscale Ag2Te precipitates and La doping. Adv
Funct Mater 2011;21:241–9.
[38] Lee MH, Rhyee J-S, Vaseem M, Hahn Y-B, Park S-D, Jin Kim H, et al. Thermoelectric properties of SrTiO3 nano-particles dispersed indium selenide bulk
composites. Appl Phys Lett 2013;102:223901.
[39] Zhang S, Zhu T, Yang S, Yu C, Zhao X. Phase compositions, nanoscale microstructures and thermoelectric properties in Ag2−ySbyTe1+y alloys with precipitated
Sb2Te3 plates. Acta Mater 2010;58:4160–9.
[40] Yang L, Chen Z-G, Dargusch MS, Zou J. High performance thermoelectric materials: progress and their applications. Adv Energy Mater 2017. 1701797.
[41] Hong M, Chen Z-G, Zou J. Fundamental and progress of Bi2Te3-based thermoelectric materials. Chinese Phys B 2018;27.
[42] Rhyee JS, Lee KH, Lee SM, Cho E, Kim SI, Lee E, et al. Peierls distortion as a route to high thermoelectric performance in In4Se3-δ crystals. Nature
2009;459:965–8.
[43] Shi X, Yang J, Salvador JR, Chi M, Cho JY, Wang H, et al. Multiple-filled skutterudites: high thermoelectric figure of merit through separately optimizing
electrical and thermal transports. J Am Chem Soc 2011;133:7837–46.
[44] Saramat A, Svensson G, Palmqvist AEC, Stiewe C, Mueller E, Platzek D, et al. Large thermoelectric figure of merit at high temperature in Czochralski-grown
clathrate Ba8Ga16Ge30. J Appl Phys 2006;99:023708.
[45] Liu W, Kim HS, Chen S, Jie Q, Lv B, Yao M, et al. N-type thermoelectric material Mg2Sn0.75Ge0.25 for high power generation. Proc Natl Acad Sci
2015;112:3269–74.
[46] Ning H, Mastrorillo GD, Grasso S, Du B, Mori T, Hu C, et al. Enhanced thermoelectric performance of porous magnesium tin silicide prepared using pressure-less
spark plasma sintering. J Mater Chem A 2015;3:17426–32.
[47] Basu R, Bhattacharya S, Bhatt R, Roy M, Ahmad S, Singh A, et al. Improved thermoelectric performance of hot pressed nanostructured n-type SiGe bulk alloys. J
Mater Chem A 2014;2:6922–30.
[48] Hong M, Chen ZG, Yang L, Zou J. BixSb2−xTe3 nanoplates with enhanced thermoelectric performance due to sufficiently decoupled electronic transport
properties and strong wide-frequency phonon scatterings. Nano Energy 2016;20:144–55.
[49] May AF, Flage-Larsen E, Snyder GJ. Electron and phonon scattering in the high-temperature thermoelectric La3Te4−zMz (M = Sb, Bi). Phys Rev B
2010;81:125205.
[50] Zhang J, Song L, Pedersen SH, Yin H, Hung LT, Iversen BB. Discovery of high-performance low-cost n-type Mg3Sb2-based thermoelectric materials with multi-
valley conduction bands. Nat Commun 2017;8:13901.
[51] Korkosz RJ, Chasapis TC, S-h Lo, Doak JW, Kim YJ, Wu C-I, et al. High ZT in p-Type (PbTe)1–2x(PbSe)x(PbS)x Thermoelectric Materials. J Am Chem Soc
2014;136:3225–37.
[52] Duong AT, Van Quang Nguyen GD, Van Thiet Duong SK, Song JY, Lee JK, Lee JE, et al. Achieving ZT = 2.2 with Bi-doped n-type SnSe single crystals. Nat
Commun 2016;7:13713.
[53] Rogl G, Sauerschnig P, Rykavets Z, Romaka V, Heinrich P, Hinterleitner B, et al. (V, Nb)-doped half Heusler alloys based on Ti, Zr, Hf NiSn with high ZT. Acta
Mater 2017;131:336–48.
[54] Fujita K, Mochida T, Nakamura K. High-temperature thermoelectric properties of NaxCoO2-δ single crystals. Jpn J Appl Phys 2001;40:4644.
[55] Brown SR, Kauzlarich SM, Gascoin F, Snyder GJ. Yb14MnSb11: new high efficiency thermoelectric material for power generation. Chem Mater 2006;18:1873–7.
[56] Schröder T, Rosenthal T, Giesbrecht N, Nentwig M, Maier S, Wang H, et al. Nanostructures in Te/Sb/Ge/Ag (TAGS) thermoelectric materials induced by phase
transitions associated with vacancy ordering. Inorg Chem 2014;53:7722–9.
[57] Girard SN, He J, Zhou X, Shoemaker D, Jaworski CM, Uher C, et al. High performance Na-doped PbTe-PbS thermoelectric materials: electronic density of states
modification and shape-controlled nanostructures. J Am Chem Soc 2011;133:16588–97.
[58] Ma Y, Hao Q, Poudel B, Lan Y, Yu B, Wang D, et al. Enhanced thermoelectric figure-of-merit in p-type nanostructured bismuth antimony tellurium alloys made
from elemental chunks. Nano Lett 2008;8:2580–4.
[59] Fu C, Zhu T, Liu Y, Xie H, Zhao X. Band engineering of high performance p-type FeNbSb based half-Heusler thermoelectric materials for figure of merit ZT > 1.
Energy Environ Sci 2015;8:216–20.
[60] Okamura C, Ueda T, Hasezaki K. Preparation of single phase β-Zn4Sb3 thermoelectric materials by mechanical grinding process. Mater Trans 2010;51:152–5.
[61] Zhao LD, Lo SH, Zhang Y, Sun H, Tan G, Uher C, et al. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature
2014;508:373–7.
[62] Rogl G, Grytsiv A, Heinrich P, Bauer E, Kumar P, Peranio N, et al. New bulk p-type skutterudites DD0.7Fe2.7Co1.3Sb12−xXx (X = Ge, Sn) reaching ZT > 1.3. Acta
Mater 2015;91:227–38.
[63] Olvera A, Moroz N, Sahoo P, Ren P, Bailey T, Page A, et al. Partial indium solubility induces chemical stability and colossal thermoelectric figure of merit in
Cu2Se. Energy Environ Sci 2017;10:1668–76.
[64] Zhang H, Talapin DV. Thermoelectric tin selenide: the beauty of simplicity. Angew Chem Int Ed 2014;53:9126–7.
[65] Sassi S, Candolfi C, Vaney JB, Ohorodniichuk V, Masschelein P, Dauscher A, et al. Assessment of the thermoelectric performance of polycrystalline p-type SnSe.
Appl Phys Lett 2014;104:212105.
[66] Popuri SR, Pollet M, Decourt R, Morrison FD, Bennett NS, Bos JWG. Large thermoelectric power factors and impact of texturing on the thermal conductivity in
polycrystalline SnSe. J Mater Chem C 2016;4:1685–91.
[67] Chen Y-X, Ge Z-H, Yin M, Feng D, Huang X-Q, Zhao W, et al. Understanding of the extremely low thermal conductivity in high-performance polycrystalline SnSe
through potassium doping. Adv Funct Mater 2016;26:6836–45.
[68] Zhao L-D, Chang C, Tan G, Kanatzidis MG. SnSe: a remarkable new thermoelectric material. Energy Environ Sci 2016;9:3044–60.
[69] Shi W, Gao M, Wei J, Gao J, Fan C, Ashalley E, et al. Tin Selenide (SnSe): growth, properties, and applications. Adv Sci 2018;1700602.
[70] Khan AA, Khan I, Ahmad I, Ali Z. Thermoelectric studies of IV–VI semiconductors for renewable energy resources. Mat Sci Semicon Proc 2016;48:85–94.
[71] Chattopadhyay T, Pannetier J, Von Schnering H. Neutron diffraction study of the structural phase transition in SnS and SnSe. J Phys Chem Solids
1986;47:879–85.
[72] Okamoto H. Se-Sn (selenium-tin). J Phase Equilib 1998;19:293.
[73] Vyas S, Pandya G, Desai C. Microhardness studies on cleavage planes of SnS and SnSe single crystals. Indian J Pure Ap Phy 1995;33:191–4.
[74] Colin R, Drowart J. Thermodynamic study of tin selenide and tin telluride using a mass spectrometer. Trans Faraday Soc 1964;60:673–83.
[75] Hirayama C, Ichikawa Y, DeRoo AM. Vapor pressures of tin selenide and tin telluride. J Phys Chem 1963;67:1039–42.
[76] Melekh B, Nb S, Sa S, Fomina T. Thermodynamic properties of compounds in tin-selenium system. Russ J Phys Chem+ 1971;45:1144.
[77] Wiedemeier H, Pultz G, Gaur U, Wunderlich B. Heat capacity measurements of SnSe and SnSe2. Thermochim Acta 1981;43:297–303.
[78] Zhao L-D, Tan G, Hao S, He J, Pei Y, Chi H, et al. Ultrahigh power factor and thermoelectric performance in hole-doped single-crystal SnSe. Science
2016;351:141–4.
[79] Peng K, Lu X, Zhan H, Hui S, Tang X, Wang G, et al. Broad temperature plateau for high ZTs in heavily doped p-type SnSe single crystals. Energy Environ Sci
2016;9:454–60.
[80] Bhatt V, Gireesan K, Pandya G. Growth and characterization of SnSe and SnSe2 single crystals. J Cryst Growth 1989;96:649–51.
[81] Li Y, Li F, Dong J, Ge Z, Kang F, He J, et al. Enhanced mid-temperature thermoelectric performance of textured SnSe polycrystals made of solvothermally
synthesized powders. J Mater Chem C 2016;4:2047–55.
[82] Han G, Popuri SR, Greer HF, Bos JW, Zhou W, Knox AR, et al. Facile surfactant-free synthesis of p-type SnSe nanoplates with exceptional thermoelectric power

338
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

factors. Angew Chem Int Ed 2016;55:6433–7.


[83] Bernardes-Silva AC, Mesquita AF, de Moura Neto E, Porto AO, de Lima GM, Ardisson JD, et al. Tin selenide synthesized by a chemical route: the effect of the
annealing conditions in the obtained phase. Solid State Commun 2005;135:677–82.
[84] Tang G, Wei W, Zhang J, Li Y, Wang X, Xu G, et al. Realizing high figure of merit in phase-separated polycrystallin Sn1-xPbxSe. J Am Chem Soc
2016;138:13647–54.
[85] Yang S, Si J, Su Q, Wu H. Enhanced thermoelectric performance of SnSe doped with layered MoS2/graphene. Mater Lett 2017;193:146–9.
[86] Bera T, Sanchela AV, Tomy C, Thakur AD. N-type SnSe1-x for thermoelectric application. Cond Mat Mtrl Sci 2016. 1601.00753.
[87] Sharma R, Chang Y. The Se−Sn (selenium-tin) system. J Phase Equilib 1986;7:68–72.
[88] Serrano-Sánchez F, Gharsallah M, Nemes NM, Mompean FJ, Martínez JL, Alonso JA. Record Seebeck coefficient and extremely low thermal conductivity in
nanostructured SnSe. Appl Phys Lett 2015;106:083902.
[89] Fu Y, Xu J, Liu G-Q, Yang J, Tan X, Liu Z, et al. Enhanced thermoelectric performance in p-type polycrystalline SnSe benefiting from texture modulation. J Mater
Chem C 2016;4:1201–7.
[90] Tyagi K, Gahtori B, Bathula S, Singh NK, Bishnoi S, Auluck S, et al. Electrical transport and mechanical properties of thermoelectric tin selenide. RSC Adv
2016;6:11562–9.
[91] Shi X, Chen Z-G, Liu W, Yang L, Hong M, Moshwan R, et al. Achieving high Figure of Merit in p-type polycrystalline Sn0.98Se via self-doping and anisotropy-
strengthening. Energy Storage Mater 2018;10:130–8.
[92] Agarwal A, Chaki SH, Lakshminarayana D. Growth and thermal studies of SnSe single crystals. Mater Lett 2007;61:5188–90.
[93] Peters MJ, McNeil LE. High-pressure Mössbauer study of SnSe. Phys Rev B 1990;41:5893–7.
[94] Pletikosić I, von Rohr F, Pervan P, Das P, Vobornik I, Cava R, et al. Band structure of a IV-VI black phosphorus analogue, the thermoelectric SnSe. Cond Mat Mtrl
Sci 2017. 1707.04289.
[95] Erdemir A. Crystal chemistry and solid lubricating properties of the monochalcogenides gallium selenide and tin selenide. Tribol T 2008;37:471–8.
[96] Abrikosov NK. Semiconducting II–VI, IV–VI, and V-VI compounds. NY: Springer; 2013.
[97] Kim S-u, Duong A-T, Cho S, Rhim SH, Kim J. A microscopic study investigating the structure of SnSe surfaces. Surf Sci 2016;651:5–9.
[98] Sist M, Zhang J, Brummerstedt Iversen B. Crystal structure and phase transition of thermoelectric SnSe. Acta Crystallogr B 2016;72:310–6.
[99] Delaire O. Phonon scattering in thermoelectrics: thermal transport, strong anharmonicity, and emergent quasiparticles. Bull Am Phys Soc 2017;62.
[100] Zhang X, Zhao L-D. Thermoelectric materials: energy conversion between heat and electricity. J Materiomics 2015;1:92–105.
[101] Heremans JP. Thermoelectricity: the ugly duckling. Nature 2014;508:327–8.
[102] Heremans JP. Thermoelectric materials: the anharmonicity blacksmith. Nat Phys 2015;11:990–1.
[103] Liu W, Yan X, Chen G, Ren Z. Recent advances in thermoelectric nanocomposites. Nano Energy 2012;1:42–56.
[104] Androulakis J, Todorov I, He J, Chung DY, Dravid V, Kanatzidis M. Thermoelectrics from abundant chemical elements: high-performance nanostructured PbSe-
PbS. J Am Chem Soc 2011;133:10920–7.
[105] Dismukes JP, Ekstrom L, Steigmeier EF, Kudman I, Beers DS. Thermal and electrical properties of heavily doped Ge-Si alloys up to 1300 K. J Appl Phys
1964;35:2899.
[106] Skomorokhov AN, Trots DM, Knapp M, Bickulova NN, Fuess H. Structural behaviour of β-Cu2−δSe (δ = 0, 0.15, 0.25) in dependence on temperature studied
by synchrotron powder diffraction. J Alloy Compd 2006;421:64–71.
[107] Morelli D, Jovovic V, Heremans J. Intrinsically minimal thermal conductivity in cubic I−V−VI2 semiconductors. Phys Rev Lett 2008;101:035901.
[108] Qin G, Yan Q-B, Qin Z, Yue S-Y, Cui H-J, Zheng Q-R, et al. Hinge-like structure induced unusual properties of black phosphorus and new strategies to improve
the thermoelectric performance. Sci Rep 2014;4:6946.
[109] Liu G, Zhou J, Wang H. Anisotropic thermal expansion of SnSe from first-principles calculations based on Grüneisen's theory. Phys Chem Chem Phys
2017;19:15187–93.
[110] Guo R, Wang X, Kuang Y, Huang B. First-principles study of anisotropic thermoelectric transport properties of IV-VI semiconductor compounds SnSe and SnS.
Phys Rev B 2015;92:115202.
[111] Popuri SR, Pollet M, Decourt R, Viciu M. Bos J-WG. Evidence for hard and soft substructures in thermoelectric SnSe. Appl Phys Lett 2017;110:253903.
[112] Skelton JM, Burton LA, Parker SC, Walsh A, Kim C-E, Soon A, et al. Anharmonicity in the high-temperature Cmcm phase of SnSe: soft modes and three-phonon
interactions. Phys Rev Lett 2016;117:075502.
[113] Lee H. A theoretical model of thermoelectric transport properties for electrons and phonons. J Electron Mater 2015;45:1115–41.
[114] Hong J, Delaire O. Electronic instability and anharmonicity in SnSe. Cond Mat Mtrl Sci 2016. 1604.07077.
[115] Li CW, Hong J, May AF, Bansal D, Chi S, Hong T, et al. Orbitally driven giant phonon anharmonicity in SnSe. Nat Phys 2015;11:1063–9.
[116] Yan J, Gorai P, Ortiz B, Miller S, Barnett SA, Mason T, et al. Material descriptors for predicting thermoelectric performance. Energy Environ Sci 2015;8:983–94.
[117] Guan X, Lu P, Wu L, Han L, Liu G, Song Y, et al. Thermoelectric properties of SnSe compound. J Alloy Compd 2015;643:116–20.
[118] Ge Z-H, Wei K, Lewis H, Martin J, Nolas GS. Bottom-up processing and low temperature transport properties of polycrystalline SnSe. J Solid State Chem
2015;225:354–8.
[119] Zhang Y, Jia X, Sun H, Sun B, Liu B, Liu H, et al. Effect of high pressure on thermoelectric performance and electronic structure of SnSe via HPHT. J Alloy
Compd 2016;667:123–9.
[120] Hong M, Chen Z-G, Yang L, Zou J. Enhancing thermoelectric performance of Bi2Te3-based nanostructures through rational structure design. Nanoscale
2016;8:8681–6.
[121] Hong M, Chen Z-G, Pei Y, Yang L, Zou J. Limit of ZT enhancement in rocksalt structured chalcogenides by band convergence. Phys Rev B 2016;94:161201.
[122] Yang L, Chen Z-G, Han G, Hong M, Huang L, Zou J. Te-Doped Cu2Se nanoplates with a high average thermoelectric figure of merit. J Mater Chem A
2016;4:9213–9.
[123] Han G, Chen Z-G, Yang L, Cheng L, Jack K, Drennan J, et al. Thermal stability and oxidation of layer-structured rhombohedral In3Se4 nanostructures. Appl Phys
Lett 2013;103:263105.
[124] Cheng L, Chen Z-G, Yang L, Han G, Xu H-Y, Snyder GJ, et al. T-shaped Bi2Te3–Te heteronanojunctions: epitaxial growth, structural modeling, and thermo-
electric properties. J Phys Chem C 2013;117:12458–64.
[125] Zhu T, Liu Y, Fu C, Heremans JP, Snyder JG, Zhao X. Compromise and synergy in high-efficiency thermoelectric materials. Adv Mater 2017;29:1605884.
[126] Duvjir G, Min T, Thi Ly T, Kim T, Duong A-T, Cho S, et al. Origin of p-type characteristics in a SnSe single crystal. Appl Phys Lett 2017;110:262106.
[127] Ortiz BR, Peng H, Lopez A, Parilla PA, Lany S, Toberer ES. Effect of extended strain fields on point defect phonon scattering in thermoelectric materials. Phys
Chem Chem Phys 2015;17:19410–23.
[128] Wei TR, Wu CF, Zhang X, Tan Q, Sun L, Pan Y, et al. Thermoelectric transport properties of pristine and Na-doped SnSe(1-x)Te(x) polycrystals. Phys Chem Chem
Phys 2015;17:30102–9.
[129] Han Y-M, Zhao J, Zhou M, Jiang X-X, Leng H-Q, Li L-F. Thermoelectric performance of SnS and SnS–SnSe solid solution. J Mater Chem A 2015;3:4555–9.
[130] Wang FQ, Zhang S, Yu J, Wang Q. Thermoelectric properties of single-layered SnSe sheet. Nanoscale 2015;7:15962–70.
[131] Wasscher J, Albers W, Haas C. Simple evaluation of the maximum thermoelectric figure of merit, with application to mixed crystals SnS1-xSex. Solid State
Electron 1963;6:261–4.
[132] Kim SI, Lee KH, Mun HA, Kim HS, Hwang SW, Roh JW, et al. Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermo-
electrics. Science 2015;348:109–14.
[133] Goldsmid HJ. Electronic refrigeration. Pion; 1986.
[134] Fan S, Zhao J, Guo J, Yan Q, Ma J, Hng HH. P-type Bi0.4Sb1.6Te3 nanocomposites with enhanced figure of merit. Appl Phys Lett 2010;96:182104.
[135] Majumdar A. Thermoelectricity in semiconductor nanostructures. Science 2004;303:777–8.
[136] Xie W, Tang X, Yan Y, Zhang Q, Tritt TM. Unique nanostructures and enhanced thermoelectric performance of melt-spun BiSbTe alloys. Appl Phys Lett

339
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

2009;94:102111.
[137] Biswas K, He J, Zhang Q, Wang G, Uher C, Dravid VP, et al. Strained endotaxial nanostructures with high thermoelectric figure of merit. Nat Chem
2011;3:160–6.
[138] Tang X, Xie W, Li H, Zhao W, Zhang Q, Niino M. Preparation and thermoelectric transport properties of high-performance p-type Bi2Te3 with layered na-
nostructure. Appl Phys Lett 2007;90:012102.
[139] Cao Y, Zhao X, Zhu T, Zhang X, Tu J. Syntheses and thermoelectric properties of Bi2Te3∕ Sb2Te3 bulk nanocomposites with laminated nanostructure. Appl Phys
Lett 2008;92:143106.
[140] Xiao F, Hangarter C, Yoo B, Rheem Y, Lee K-H, Myung NV. Recent progress in electrodeposition of thermoelectric thin films and nanostructures. Electrochim
Acta 2008;53:8103–17.
[141] Tang G, Wen Q, Yang T, Cao Y, Wei W, Wang Z, et al. Rock-salt-type nanoprecipitates lead to high thermoelectric performance in undoped polycrystalline SnSe.
RSC Adv 2017;7:8258–63.
[142] Zhang Y, Hao S, Zhao L-D, Wolverton C, Zeng Z. Pressure induced thermoelectric enhancement in SnSe crystals. J Mater Chem A 2016;4:12073–9.
[143] Yan J, Ke F, Liu C, Wang L, Wang Q, Zhang J, et al. Pressure-driven semiconducting-semimetallic transition in SnSe. Phys Chem Chem Phys 2016;18:5012–8.
[144] Loa I, Husband RJ, Downie RA, Popuri SR, Bos JW. Structural changes in thermoelectric SnSe at high pressures. J Phys-Condens Mat 2015;27:072202.
[145] Michielon de Souza S, Ordozgoith da Frota H, Trichês DM, Ghosh A, Chaudhuri P, dos Santos Silva, et al. Pressure-induced polymorphism in nanostructured
SnSe. J Appl Crystallogr 2016;49:213–21.
[146] Timofeev YA, Vinogradov B, Begoulev V. Superconductivity of tin selenide at pressures up to 70 GPa. Phys Solid State + 1997;39:207.
[147] Rogl G, Setman D, Schafler E, Horky J, Kerber M, Zehetbauer M, et al. High-pressure torsion, a new processing route for thermoelectrics of high ZTs by means of
severe plastic deformation. Acta Mater 2012;60:2146–57.
[148] Yu H, Dai S, Chen Y. Enhanced power factor via the control of structural phase transition in SnSe. Sci Rep 2016;6:26193.
[149] Li G, Aydemir U, Wood M, Goddard III WA, Zhai P, Zhang Q, et al. Ideal strength and deformation mechanism in high-efficiency thermoelectric SnSe. Chem
Mater 2017;29:2382–9.
[150] Fu J, Su X, Xie H, Yan Y, Liu W, You Y, et al. Understanding the combustion process for the synthesis of mechanically robust SnSe thermoelectrics. Nano Energy
2018;44:53–62.
[151] Anstis G, Chantikul P, Lawn BR, Marshall D. A critical evaluation of indentation techniques for measuring fracture toughness: I, direct crack measurements. J
Am Ceram Soc 1981;64:533–8.
[152] Ren F, Hall BD, Ni JE, Case ED, Sootsman J, Kanatzidis MG, et al. Mechanical characterization of PbTe-based thermoelectric materials. MRS Online Proceedings
Library Archive 2007;1044.
[153] Zhao L-D, Zhang B-P, Li J-F, Zhou M, Liu W-S, Liu J. Thermoelectric and mechanical properties of nano-SiC-dispersed Bi2Te3 fabricated by mechanical alloying
and spark plasma sintering. J Alloy Compd 2008;455:259–64.
[154] Tyagi K, Gahtori B, Bathula S, Jayasimhadri M, Sharma S, Singh NK, et al. Crystal structure and mechanical properties of spark plasma sintered Cu2Se: an
efficient photovoltaic and thermoelectric material. Solid State Commun 2015;207:21–5.
[155] Ngan PH, Christensen DV, Snyder GJ, Hung LT, Linderoth S, Nong NV, et al. Towards high efficiency segmented thermoelectric unicouples. Phys Status Solidi A
2014;211:9–17.
[156] Bartholomé K, Balke B, Zuckermann D, Köhne M, Müller M, Tarantik K, et al. Thermoelectric modules based on half-Heusler materials produced in large
quantities. J Electron Mater 2014;43:1775–81.
[157] Ren F, Wang H, Menchhofer PA, Kiggans JO. Thermoelectric and mechanical properties of multi-walled carbon nanotube doped Bi0.4Sb1.6Te3 thermoelectric
material. Appl Phys Lett 2013;103:221907.
[158] Salvador JR, Yang J, Wereszczak AA, Wang H, Cho JY. Temperature dependent tensile fracture stress of n-and p-type filled-skutterudite materials. Sci Adv
Mater 2011;3:577–86.
[159] Salvador J, Yang J, Shi X, Wang H, Wereszczak A, Kong H, et al. Transport and mechanical properties of Yb-filled skutterudites. Philos Mag 2009;89:1517–34.
[160] Badrinarayanan S, Mandale A, Gunjikar V, Sinha A. Mechanism of high-temperature oxidation of tin selenide. J Mater Sci 1986;21:3333–8.
[161] Li Y, He B, Heremans JP, Zhao J-C. High-temperature oxidation behavior of thermoelectric SnSe. J Alloy Compd 2016;669:224–31.
[162] Zhang B, Peng K, Sha X, Li A, Zhou X, Chen Y, et al. A second amorphous layer underneath surface oxide. Microsc Microanal 2017;23:173–8.
[163] Smulko Janusz M, Ederth J, Li Yingfeng, Kish Laszlo B, Kennedy Marcus K, Kruis Frank E. Gas sensing by thermoelectric voltage fluctuations in SnO2
nanoparticle films. Sensor Actuat B-Chem 2005;106:708–12.
[164] Yanagiya S, Nong NV, Xu J, Sonne M, Pryds N. Thermoelectric properties of SnO2 ceramics doped with Sb and Zn. J Electron Mater 2011;40:674–7.
[165] Ambrosini A, Palmer G, Maignan A, Poeppelmeier KR, Lane M, Brazis P, et al. Variable-temperature electrical measurements of zinc oxide/tin oxide-cosub-
stituted indium oxide. Chem Mater 2002;14:52–7.
[166] Ahmed SF, Khan S, Ghosh PK, Mitra MK, Chattopadhyay KK. Effect of Al doping on the conductivity type inversion and electro-optical properties of SnO2 thin
films synthesized by sol-gel technique. J Sol-gel Sci Techn 2006;39:241–7.
[167] Yu JG, Yue A, Stafsudd O. Growth and electronic properties of the SnSe semiconductor. J Cryst Growth 1981;54:248–52.
[168] Guo S-D. Thermoelectric properties of orthorhombic group IV-VI monolayers from the first-principles calculations. J Appl Phys 2017;121:034302.
[169] Feng Shi Q, Li Yan Y, Xu Wang Y. Electronic structure and thermoelectric performance of Zintl compound Sr3GaSb3: a first-principles study. Appl Phys Lett
2014;104:012104.
[170] Yan YL, Wang YX. Electronic structure and thermoelectric properties of In32−xGexO48 (x = 0, 1, 2, and 3) at low temperature. Appl Phys Lett 2010;97:252106.
[171] Gudelli VK, Kanchana V, Appalakondaiah S, Vaitheeswaran G, Valsakumar M. Phase stability and thermoelectric properties of the mineral FeS2: an ab initio
study. J Phys Chem C 2013;117:21120–31.
[172] Wang C, Wang Y, Zhang G, Peng C. Electronic structure and thermoelectric properties of ZnO single-walled nanotubes and nanowires. J Phys Chem C
2013;117:21037–42.
[173] Balout H, Boulet P, Record M-C. Thermoelectric properties of Mg2Si thin films by computational approaches. J Phys Chem C 2014;118:19635–45.
[174] Si HG, Wang YX, Yan YL, Zhang GB. Structural, electronic, and thermoelectric properties of InSe nanotubes: first-principles calculations. J Phys Chem C
2012;116:3956–61.
[175] Zou D, Xie S, Liu Y, Lin J, Li J. First-principles study of thermoelectric and lattice vibrational properties of chalcopyrite CuGaTe2. J Alloy Compd
2013;570:150–5.
[176] Yadav MK, Sanyal B. First principles study of thermoelectric properties of Li-based half-Heusler alloys. J Alloy Compd 2015;622:388–93.
[177] Jayaraman A, Molli M, Kamisetti V. Effect of sulfur doping on thermoelectric properties of tin selenide – a first principles study. 2015;1667:110046.
[178] Do DC, Rhim S, Lee J-H, Hong SC. The electrical and thermoelectric properties of isoelectronic doping in SnSe: a first principles study. Bull Am Phys Soc
2017;62.
[179] Kutorasinski K, Wiendlocha B, Kaprzyk S, Tobola J. Electronic structure and thermoelectric properties of n- and p-type SnSe from first-principles calculations.
Phys Rev B 2015;91:205201.
[180] Xu B, Zhang J, Yu G, Ma S, Wang Y, Yi L. Comparative study of electronic structure and thermoelectric properties of SnSe for Pnma and Cmcm phase. J Electron
Mater 2016;45:5232–7.
[181] Shi G, Kioupakis E. Quasiparticle band structures and thermoelectric transport properties of p-type SnSe. J Appl Phys 2015;117:065103.
[182] Cutler M, Mott NF. Observation of Anderson localization in an electron gas. Phys Rev 1969;181:1336.
[183] Cuong DD, Rhim SH, Lee J-H, Hong SC. Strain effect on electronic structure and thermoelectric properties of orthorhombic SnSe: a first principles study. AIP
Advances 2015;5:117147.
[184] Denton AR, Ashcroft NW. Vegard’s law. Phys Rev A 1991;43:3161.
[185] Koh YK, Cao Y, Cahill DG, Jena D. Heat-transport mechanisms in superlattices. Adv Funct Mater 2009;19:610–5.

340
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[186] Callaway J. Model for lattice thermal conductivity at low temperatures. Phys Rev 1959;113:1046.
[187] Abeles B. Lattice thermal conductivity of disordered semiconductor alloys at high temperatures. Phys Rev 1963;131:1906.
[188] Suzuki Y, Nakamura H. A supercell approach to the doping effect on the thermoelectric properties of SnSe. Phys Chem Chem Phys 2015;17:29647–54.
[189] Mori H, Usui H, Ochi M, Kuroki K. Temperature-and doping-dependent roles of valleys in thermoelectric performance of SnSe: a first-principles study. Phys Rev
B 2017;96:085113.
[190] González-Romero RL, Meléndez JJ. Variation of the zT factor of SnSe with doping: a first-principles study. J Alloy Compd 2018;732:536–46.
[191] Kim I-H, Choi S-M, Seo W-S, Cheong D-I, Kang H. Synthesis and thermoelectric properties of Cu-dispersed Bi2Te2.7Se0.3. J Ceram Process Res 2012;13:170.
[192] Barczak S, Downie RA, Popuri SR, Decourt R, Pollet M, Bos J. Thermoelectric properties of Fe and Al double substituted MnSiγ (γ ∼ 1.73). J Solid State Chem
2015;227:55–9.
[193] Maji P, Makongo JP, Zhou X, Chi H, Uher C, Poudeu PF. Thermoelectric performance of nanostructured p-type Zr0.5Hf0.5Co0.4Rh0.6Sb1−xSnx half-Heusler alloys.
J Solid State Chem 2013;202:70–6.
[194] Ding G, Gao G, Yao K. High-efficient thermoelectric materials: the case of orthorhombic IV-VI compounds. Sci Rep 2015;5:9567.
[195] Carrete J, Mingo N, Curtarolo S. Low thermal conductivity and triaxial phononic anisotropy of SnSe. Appl Phys Lett 2014;105:101907.
[196] Hong AJ, Li L, Zhu HX, Yan ZB, Liu JM, Ren ZF. Optimizing the thermoelectric performance of low-temperature SnSe compounds by electronic structure design.
J Mater Chem A 2015;3:13365–70.
[197] González-Romero RL, Antonelli A, Meléndez JJ. Insights into the thermoelectric properties of SnSe from ab initio calculations. Phys Chem Chem Phys
2017;19:12804–15.
[198] Dewandre A, Hellman O, Bhattacharya S, Romero AH, Madsen GK, Verstraete MJ. Two-step phase transition in snse and the origins of its high power factor
from first principles. Phys Rev Lett 2016;117:276601.
[199] Wang Y, Hu Y-J, Shang S-L, Zhou B-C, Liu Z-K, Chen L-Q. First-principles thermodynamic theory of seebeck coefficients. Cond Mat Mtrl Sci 2017. 1708.01591.
[200] Li C, Guo D, Li K, Shao B, Chen D, Ma Y, et al. Excellent thermoelectricity performance of p-type SnSe along b-axis. Phys B: Condens Matter 2018;530:264–9.
[201] González-Romero RL, Antonelli A, Meléndez JJ. ZT-factor enhancement in SnSe: predictions from first principles calculations. Cond Mat Mtrl Sci 2016.
1612.05967.
[202] Blaha P, Schwarz K, Madsen G, Kvasnicka D, Luitz J. An augmented plane wave+ local orbitals program for calculating crystal properties. BPSKM GKH 2001.
ISBN 3-9501031-1-2.
[203] Madsen GK, Singh DJ. A code for calculating band-structure dependent quantities. Comput Phys Commun 2006;175:67–71.
[204] Yang J, Zhang G, Yang G, Wang C, Wang YX. Outstanding thermoelectric performances for both p- and n-type SnSe from first-principles study. J Alloy Compd
2015;644:615–20.
[205] Rehman SU, Butt FK, Hayat F, Haq BU, Tariq Z, Aleem F, et al. An insight into a novel cubic phase SnSe for prospective applications in optoelectronics and clean
energy devices. J Alloy Compd 2018;733:22–32.
[206] Butt FK, Haq BU, ur Rehman S, Ahmed R, Cao C, AlFaifi S. Investigation of thermoelectric properties of novel cubic phase SnSe: a promising material for
thermoelectric applications. J Alloy Compd 2017;715:438–44.
[207] ur Rehman S, Butt FK, Tariq Z, Hayat F, Gilani R, Aleem F. Pressure induced structural and optical properties of cubic phase SnSe: an investigation for the
infrared/mid-infrared optoelectronic devices. J Alloy Compd 2017;695:194–201.
[208] Lin C, Cheng W-D, Guo ZX, Chai G, Zhang H. Exceptional thermoelectric performance of a “Star-Like” SnSe nanotube with ultra-low thermal conductivity and a
high power factor. Phys Chem Chem Phys 2017;19:23247–53.
[209] Yağmurcukardeş M, Peeters FM, Senger RT, Şahin H. Nanoribbons: from fundamentals to state-of-the-art applications. Appl Phys Rev 2016;3:041302.
[210] Tyagi K, Waters K, Wang G, Gahtori B, Haranath D, Pandey R. Thermoelectric properties of SnSe nanoribbons: a theoretical aspect. Mater Res Exp
2016;3:035013.
[211] Bridgman PW. Certain physical properties of single crystals of tungsten, antimony, bismuth, tellurium, cadmium, zinc, and tin. In: Proceedings of the American
academy of arts and sciences: JSTOR; 1925. p. 305–83.
[212] Bridgman PW. Crystals and their manufacture. Google Patents; 1931.
[213] Stockbarger DC. The production of large single crystals of lithium fluoride. Rev Sci Instrum 1936;7:133.
[214] Agarwal M, Patel P, Patel J, Kshatriya J. Electron microscopy of layered single crystals grown by direct vapour transport method. B Mater Sci 1984;6:549–67.
[215] Agarwal A, Patel P, Lakshminarayana D. Single crystal growth of layered tin monoselenide semiconductor using a direct vapour transport technique. J Cryst
Growth 1994;142:344–8.
[216] Nariya B, Dasadia A, Bhayani M, Patel A, Jani A. Electrical transport properties of SnS and SnSe single crystals grown by direct vapour transport technique.
Chalcogenide Lett 2009;6:549–54.
[217] Nguyen TMH, Duong AT, Duvjir G, Trinh TL, Nguyen VQ, Kim J, et al. Thermoelectric properties of SnSe1-xSx (0 < x < 1) single crystals. Bull Am Phys Soc
2017;62.
[218] Ly TT, Duvjir G, Min T, Byun J, Kim T, Saad MM, et al. Atomistic study of the alloying behavior of crystalline SnSe1−xSx. Phys Chem Chem Phys
2017;19:21648–54.
[219] Lyell C. A manual of elementary geology. Murray; 1855.
[220] O'Donoghue M. A guide to Man-made Gemstones. Van Nostrand Reinhold Company; 1983.
[221] Byrappa K, Yoshimura M. Handbook of hydrothermal technology. William Andrew; 2012.
[222] Byrappa K, Yoshimura M. Handbook of hydrothermal technology a technology for crystal growth and materials processing. Toronto: Will‰ Andrew Publishing;
2001.
[223] Choi J, Lee H-W, Kim B-S, Park H, Choi S, Hong S, et al. Magnetic and transport properties of Mn-doped Bi2Se3 and Sb2Se3. J Magn Magn Mater
2006;304:e164–6.
[224] Li B, Xie Y, Huang J, Qian Y. Solvothermal route to tin monoselenide bulk single crystal with different morphologies. Inorg Chem 2000;39:2061–4.
[225] Born M. Thermodynamics of crystals and melting. J Chem Phys 1939;7:591–603.
[226] Glaser MA, Clark NA. Melting and liquid structure in two dimensions. Adv Chem Phys 1993;83:543–709.
[227] Christenson HK. Confinement effects on freezing and melting. J Phys-Condens Mater 2001;13:R95.
[228] Schmalzried H. Solid-State Reactions. Angew Chem Int Ed Engl 1963;2:251–4.
[229] Schmalzried H. Solid-state reactions. Treatise on solid state chemistry. Springer; 1976. p. 233–79.
[230] Schmalzried H. Solid state reactions. Vch Pub; 1981.
[231] Agarwal A, Vashi M, Lakshminarayana D, Batra N. Electrical resistivity anisotropy in layered p-SnSe single crystals. J Mater Sci-Mater El 2000;11:67–71.
[232] Hong M, Chen Z-G, Yang L, Chasapis TC, Kang SD, Zou Y, et al. Enhancing the thermoelectric performance of SnSe1−xTex nanoplates through band engineering.
J Mater Chem A 2017;5:10713–21.
[233] Guo J, Jian J, Liu J, Cao B, Lei R, Zhang Z, et al. Synthesis of SnSe nanobelts and the enhanced thermoelectric performance in its hot-pressed bulk composite.
Nano Energy 2017;38:569–75.
[234] Seward T. Metal complex formation in aqueous solutions at elevated temperatures and pressures. Phys Chem Earth 1981;13:113–32.
[235] Han G, Popuri SR, Greer HF, Llin LF, Bos JWG, Zhou W, et al. Chlorine-Enabled Electron Doping in Solution-Synthesized SnSe Thermoelectric Nanomaterials.
Adv Energy Mater 2017;7:1602328.
[236] Kim J, Duvjir G, Ly TT, Min T, Kim T, Kim SH, et al. STM study on the surface structures and defects of SnSe. Bull Am Phys Soc 2017;62.
[237] Wu D, Wu L, He D, Zhao L-D, Li W, Wu M, et al. Direct observation of vast off-stoichiometric defects in single crystalline SnSe. Nano Energy 2017;35:321–30.
[238] Shen J, Fu C, Liu Y, Zhao X, Zhu T. Enhancing thermoelectric performance of FeNbSb half-Heusler compound by Hf-Ti dual-doping. Energy Storage Mater
2017;10:69–74.
[239] Zhao L-D, Dravid VP, Kanatzidis MG. The panoscopic approach to high performance thermoelectrics. Energy Environ Sci 2014;7:251–68.

341
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[240] Thesberg M, Kosina H, Neophytou N. On the Lorenz number of multiband materials. Phys Rev B 2017;95:125206.
[241] Kuroki K, Arita R. “Pudding mold” band drives large thermopower in NaxCoO2. J Phys Soc Jpn 2007;76:083707.
[242] Jin M, Shao H, Hu H, Li D, Xu J, Liu G, et al. Single crystal growth of Sn0.97Ag0.03Se by a novel horizontal Bridgman method and its thermoelectric properties. J
Cryst Growth 2017;460:112–6.
[243] Lu Q, Wu M, Wu D, Chang C, Guo Y, Zhou C, et al. Unexpected large hole effective masses in SnSe revealed by angle-resolved photoemission spectroscopy. Phys
Rev Lett 2017;119:116401.
[244] Cho S, Nguyen VQ, Duvjir G, Duong VT, Kwon S, Song J, et al. The achievement of high ZT in n-type SnSe single crystal. Bull Am Phys Soc 2017;62.
[245] Wei P-C, Bhattacharya S, He J, Neeleshwar S, Podila R, Chen Y, et al. The intrinsic thermal conductivity of SnSe. Nature 2016;539:E1–2.
[246] Ibrahim D, Vaney J-B, Sassi S, Candolfi C, Ohorodniichuk V, Levinsky P, et al. Reinvestigation of the thermal properties of single-crystalline SnSe. Appl Phys
Lett 2017;110:032103.
[247] Albers W, Haas C, Ober H, Schodder G, Wasscher J. Preparation and properties of mixed crystals SnS(1−x)Sex. J Phys Chem Solids 1962;23:215–20.
[248] Park SH, Jin Y, Ahn K, Chung I, Yoo C-Y. Ag/Ni metallization bilayer: a functional layer for highly efficient polycrystalline SnSe thermoelectric modules. J
Electron Mater 2017;46:848–55.
[249] Li Y, Shi X, Ren D, Chen J, Chen L. Investigation of the anisotropic thermoelectric properties of oriented polycrystalline SnSe. Energies 2015;8:6275–85.
[250] Morales Ferreiro JO, Diaz-Droguett DE, Celentano D, Reparaz JS, Sotomayor Torres CM, Ganguli S, et al. Effect of the annealing on the power factor of un-
doped cold-pressed SnSe. Appl Therm Eng 2016;111:1426–32.
[251] Fanciulli C, Coduri M, Boldrini S, Abedi H, Tomasi C, Famengo A, et al. Structural texture induced in snse thermoelectric compound via open die pressing. J
Nanosci Nanotechnol 2017;17:1571–8.
[252] Li D, Li J, Qin X, Zhang J, Song C, Wang L, et al. Thermoelectric performance for SnSe hot-pressed at different temperature. J Electron Mater 2017;46:79–84.
[253] Boulet B, Lalli G, Ajersch M. Modeling and control of an electric arc furnace. In: American control conference: IEEE; 2003. p. 3060–4.
[254] Serrano-Sánchez F, Gharsallah M, Bermúdez J, Carrascoso F, Nemes NM, Dura OJ, et al. Nanostructured state-of-the-art thermoelectric materials prepared by
straight-forward arc-melting method. In: Skipidarov S, Nikitin M, editors. Thermoelectrics for power generation - a look at trends in the technology. rijeka:
InTech; 2016. p. Ch. 08.
[255] Van Vlack LH. Elements of materials science and engineering. Addison-Wesley; 1989.
[256] Suryanarayana C. Mechanical alloying and milling. Prog Mater Sci 2001;46:1–184.
[257] Wu CF, Wei TR, Li JF. Electrical and thermal transport properties of Pb(1-x)Sn(x)Se solid solution thermoelectric materials. Phys Chem Chem Phys
2015;17:13006–12.
[258] Chere EK, Zhang Q, Dahal K, Cao F, Mao J, Ren Z. Studies on thermoelectric figure of merit of Na-doped p-type polycrystalline SnSe. J Mater Chem A
2016;4:1848–54.
[259] Zhang Q, Chere EK, Sun J, Cao F, Dahal K, Chen S, et al. Studies on thermoelectric properties of n-type polycrystalline SnSe1-xSx by iodine doping. Adv Energy
Mater 2015;5:1500360.
[260] Ju H, Kim J. Effect of SiC ceramics on thermoelectric properties of SiC/SnSe composites for solid-state thermoelectric applications. Ceram Int 2016;42:9550–6.
[261] German RM. AZ of powder metallurgy. Elsevier Science Limited; 2005.
[262] Li JC, Li D, Qin XY, Zhang J. Enhanced thermoelectric performance of p-type SnSe doped with Zn. Scripta Mater 2017;126:6–10.
[263] Sairam K, Sonber J, Murthy TC, Subramanian C, Fotedar R, Nanekar P, et al. Influence of spark plasma sintering parameters on densification and mechanical
properties of boron carbide. Int J Refractory Metals Hard Mater 2014;42:185–92.
[264] Boscher ND, Carmalt CJ, Palgrave RG, Parkin IP. Atmospheric pressure chemical vapour deposition of SnSe and SnSe2 thin films on glass. Thin Solid Films
2008;516:4750–7.
[265] Han Q, Zhu Y, Wang X, Ding W. Room temperature growth of SnSe nanorods from aqueous solution. J Mater Sci 2004;39:4643–6.
[266] Ananthi K, Thilakavathy K, Muthukumarasamy N, Dhanapandian S, Murali KR. Properties of pulse plated SnSe films. J Mater Sci-Mater El 2011;23:1338–41.
[267] Chen S, Cai K, Zhao W. The effect of Te doping on the electronic structure and thermoelectric properties of SnSe. Phys B: Condens Matter 2012;407:4154–9.
[268] Chen C-L, Wang H, Chen Y-Y, Day T, Snyder GJ. Thermoelectric properties of p-type polycrystalline SnSe doped with Ag. J Mater Chem A 2014;2:11171.
[269] Singh NK, Bathula S, Gahtori B, Tyagi K, Haranath D, Dhar A. The effect of doping on thermoelectric performance of p-type SnSe: promising thermoelectric
material. J Alloy Compd 2016;668:152–8.
[270] Ge Z-H, Song D, Chong X, Zheng F, Jin L, Qian X, et al. Boosting the thermoelectric performance of (Na, K) co-doped polycrystalline SnSe by synergistic
tailoring of the band structure and atomic-scale defect phonon scattering. J Am Chem Soc 2017;139:9714–20.
[271] Sun Y, Zhong Z, Shirakawa T, Franchini C, Li D, Li Y, et al. Rocksalt SnS and SnSe: native topological crystalline insulators. Phys Rev B 2013;88:235122.
[272] Mariano A, Chopra K. Polymorphism in some IV-VI compounds induced by high pressure and thin-film epitaxial growth. Appl Phys Lett 1967;10:282–4.
[273] Wang Z, Wang J, Zang Y, Zhang Q, Shi JA, Jiang T, et al. Molecular beam epitaxy-Grown SnSe in the rock-salt structure: an artificial topological crystalline
insulator material. Adv Mater 2015;27:4150–4.
[274] Leng H-Q, Zhou M, Zhao J, Han Y-M, Li L-F. The thermoelectric performance of anisotropic SnSe doped with Na. RSC Adv 2016;6:9112–6.
[275] Zhang L, Wang J, Sun Q, Qin P, Cheng Z, Ge Z, et al. Three-stage inter-orthorhombic evolution and high thermoelectric performance in Ag-doped nanolaminar
SnSe polycrystals. Adv Energy Mater 2017;7:1700573.
[276] Guo H, Xin H, Qin X, Zhang J, Li D, Li Y, et al. Enhanced thermoelectric performance of highly oriented polycrystalline SnSe based composites incorporated
with SnTe nanoinclusions. J Alloy Compd 2016;689:87–93.
[277] Leng H, Zhou M, Zhao J, Han Y, Li L. Optimization of thermoelectric performance of anisotropic AgxSn1−xSe compounds. J Electron Mater 2015;45:527–34.
[278] Kucek V, Plechacek T, Janicek P, Ruleova P, Benes L, Navratil J, et al. Thermoelectric properties of Tl-Doped SnSe: a hint of phononic structure. J Electron
Mater 2016;45:2943–9.
[279] Kim JH, Oh S, Kim YM, So HS, Lee H, Rhyee J-S, et al. Indium substitution effect on thermoelectric and optical properties of Sn1−xInxSe compounds. J Alloy
Compd 2016;682:785–90.
[280] Quang NV, Tuan DA, Thiet DV, Minh Hai NT, Duvjir G, Ly TT, et al. Optimization of p-type SnSe thermoelectric performance by controlling vacancies. Bull Am
Phys Soc 2017;62.
[281] Wang X, Xu J, Liu G, Fu Y, Liu Z, Tan X, et al. Optimization of thermoelectric properties in n-type SnSe doped with BiCl3. Appl Phys Lett 2016;108:083902.
[282] Li Q, Zhang L, Yin J, Sheng Z, Chu X, Wang F, et al. Study on the thermoelectric performance of polycrystal SnSe with Se vacancies. J Alloy Compd
2018;745:513–8.
[283] Wei TR, Tan G, Zhang X, Wu CF, Li JF, Dravid VP, et al. Distinct impact of alkali-ion doping on electrical transport properties of thermoelectric p-type
polycrystalline SnSe. J Am Chem Soc 2016;138:8875–82.
[284] Wei T-R, Tan G, Wu C-F, Chang C, Zhao L-D, Li J-F, et al. Thermoelectric transport properties of polycrystalline SnSe alloyed with PbSe. Appl Phys Lett
2017;110:053901.
[285] Cai B, Li J, Sun H, Zhao P, Yu F, Zhang L, et al. Sodium doped polycrystalline SnSe: high pressure synthesis and thermoelectric properties. J Alloy Compd
2017;727:1014–9.
[286] Lee YK, Ahn K, Cha J, Zhou C, Kim HS, Choi G, et al. Enhancing p-type thermoelectric performances of polycrystalline SnSe via tuning phase transition
temperature. J Am Chem Soc 2017;139:10887–96.
[287] Gao J, Xu G. Thermoelectric performance of polycrystalline Sn1-xCuxSe (x = 0–0.03) prepared by high pressure method. Intermetallics 2017;89:40–5.
[288] Chien C-H, Chang C-C, Chen C-L, Tseng C-M, Wu Y-R, Wu M-K, et al. Facile chemical synthesis and enhanced thermoelectric properties of Ag doped SnSe
nanocrystals. RSC Adv 2017;7:34300–6.
[289] Gharsallah M, Serrano-Sanchez F, Nemes NM, Mompean FJ, Martinez JL, Fernandez-Diaz MT, et al. Giant seebeck effect in Ge-doped SnSe. Sci Rep
2016;6:26774.
[290] Fu Y, Xu J, Liu G-Q, Tan X, Liu Z, Wang X, et al. Study on thermoelectric properties of polycrystalline SnSe by Ge doping. J Electron Mater 2017;46:3182–6.

342
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[291] Gao J, Shao Z, Xu G. Thermoelectric transport properties of Sn1-xGexSe (x = 0–0.03) prepared by melting synthesis method. Int J Appl Ceram Tec; 14:963–8.
[292] Gao J, Zhu H, Mao T, Zhang L, Di J, Xu G. The effect of Sm doping on the transport and thermoelectric properties of SnSe. Mater Res Bull 2017;93:366–72.
[293] Li F, Wang W, Ge Z-H, Zheng Z, Luo J, Fan P, et al. Enhanced thermoelectric properties of polycrystalline SnSe via LaCl3 doping. Materials 2018;11:203.
[294] Li D, Tan X, Xu J, Liu G, Jin M, Shao H, et al. Enhanced thermoelectric performance in n-type polycrystalline SnSe by PbBr 2 doping. RSC Adv
2017;7:17906–12.
[295] Li F, Wang W, Qiu X, Zheng Z-h, Fan P, J-t Luo, et al. Optimization of thermoelectric properties of n-type Ti, Pb co-doped SnSe. Inorg Chem Front
2017;4:1721–9.
[296] Gong Y, Chang C, Wei W, Liu J, Xiong W, Chai S, et al. Extremely low thermal conductivity and enhanced thermoelectric performance of polycrystalline SnSe by
Cu doping. Scripta Mater 2018;147:74–8.
[297] Ju H, Park D, Kim J. Organic acid-assisted chemical transformation of SnSe bulk crystals into Se nanowires and the influence on thermoelectric behaviors. J
Alloy Compd 2018;732:436–42.
[298] Li J, Li D, Xu W, Qin X, Li Y, Zhang J. Enhanced thermoelectric performance of SnSe based composites with carbon black nanoinclusions. Appl Phys Lett
2016;109:173902.
[299] Wei T-R, Li Z, Sun F-H, Pan Y, Wu C-F, Farooq MU, et al. Thermoelectric SnS and SnS-SnSe solid solutions prepared by mechanical alloying and spark plasma
sintering: anisotropic thermoelectric properties. Sci Rep 2017;7:43262.
[300] Huang X-Q, Chen Y-X, Yin M, Feng D, He J. Origin of the enhancement in transport properties on polycrystalline SnSe with compositing two-dimensional
material MoSe2. Nanotechnology 2017;28:105708.
[301] Li D, Li J, Qin X, Zhang J, Xin H, Song C, et al. Enhanced thermoelectric performance in SnSe based composites with PbTe nanoinclusions. Energy
2016;116:861–6.
[302] Lin C-C, Lydia R, Yoon JH, Lee HS, Rhyee JS. Extremely low lattice thermal conductivity and point defect scattering of phonons in Ag-doped (SnSe)1− x(SnS)x
compounds. Chem Mater 2017;29:5344–52.
[303] Liu FS, Huang MJ, Gong ZN, Ao WQ, Li Y, Li JQ. Enhancing the thermoelectric performance of β-Cu2Se by incorporating SnSe. J Alloy Compd
2015;651:648–54.
[304] Siol S, Holder A, Ortiz BR, Parilla PA, Toberer E, Lany S, et al. Solubility limits in quaternary SnTe-based alloys. RSC Adv 2017;7:24747–53.
[305] Chu F, Zhang Q, Zhou Z, Hou D, Wang L, Jiang W. Enhanced thermoelectric and mechanical properties of Na-doped polycrystalline SnSe thermoelectric
materials via CNTs dispersion. J Alloy Compd 2018;741:756–64.
[306] Li Z, Sun F-H, Tang H, Dong J-F, Li J-F. Enhanced thermoelectric properties of p-type SnS0.2Se0.8 solid solution doped with Ag. J Alloy Compd
2018;745:172–8.
[307] Hicks L, Dresselhaus M. Thermoelectric figure of merit of a one-dimensional conductor. Phys Rev B 1993;47:16631.
[308] Hicks L, Harman T, Dresselhaus M. Use of quantum-well superlattices to obtain a high figure of merit from nonconventional thermoelectric materials. Appl Phys
Lett 1993;63:3230–2.
[309] Rao TS, Samantharay B, Chaudhuri A. Structural characterization of tin selenide thin films. J Mater Sci Lett 1985;4:743–5.
[310] Zhou Y, Zhao L-D. Promising thermoelectric bulk materials with 2D structures. Adv Mater:1702676.
[311] Tayari V, Senkovskiy B, Rybkovskiy D, Ehlen N, Fedorov A, Chen C-Y, et al. Quasi-two-dimensional thermoelectricity in SnSe. Phys Rev B 2018;97:045424.
[312] Singh AK, Mathew K, Zhuang HL, Hennig RG. Computational screening of 2D materials for photocatalysis. J Phys Chem Lett 2015;6:1087–98.
[313] Zhang LC, Qin G, Fang WZ, Cui HJ, Zheng QR, Yan QB, et al. Tinselenidene: a two-dimensional auxetic material with ultralow lattice thermal conductivity and
ultrahigh hole mobility. Sci Rep 2016;6:19830.
[314] Liu F, Hu L, Rao R, Darroudi T, Lee P-C, Bhattacharya S, et al. Electron and phonon transport in SnSe 2D nanoplatelets. Bull Am Phys Soc 2017;62.
[315] Wang X, Jones AM, Seyler KL, Tran V, Jia Y, Zhao H, et al. Highly anisotropic and robust excitons in monolayer black phosphorus. Nat Nano 2015;10:517–21.
[316] Novoselov KS, Geim AK, Morozov S, Jiang D, Katsnelson M, Grigorieva I, et al. Two-dimensional gas of massless Dirac fermions in graphene. Nature
2005;438:197–200.
[317] Shafique A, Shin Y-H. Thermoelectric and phonon transport properties of two-dimensional IV–VI compounds. Sci Rep 2017;7:506.
[318] Qiao J, Kong X, Hu Z-X, Yang F, Ji W. High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus. Nat Commun 2014;5:4475.
[319] Xia F, Wang H, Jia Y. Rediscovering black phosphorus as an anisotropic layered material for optoelectronics and electronics. Nat Commun 2014;5:4458.
[320] Baroni S, De Gironcoli S, Dal Corso A, Giannozzi P. Phonons and related crystal properties from density-functional perturbation theory. Rev Mod Phys
2001;73:515.
[321] Sirikumara H, Jayasekera T. Bonding directionality matters: direct-indirect transition in few-layer SnSe. Bull Am Phys Soc 2017;62.
[322] Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett 1996;77:3865.
[323] Heyd J, Scuseria GE. Efficient hybrid density functional calculations in solids: assessment of the Heyd–Scuseria–Ernzerhof screened Coulomb hybrid functional.
J Chem Phys 2004;121:1187–92.
[324] Mori H, Usui H, Ochi M, Kuroki K. First-principles study on the high thermoelectric efficiency originating from“pudding-mold”bands in n-and p-type SnSe. Bull
Am Phys Soc 2017;62.
[325] Wang Z, Fan C, Shen Z, Hua C, Hu Y, Sheng F, et al. Defects controlled hole doping and multi-valley transport in SnSe single crystals. Nat Commun 2018;9:47.
[326] Shi H, Parker D, Du M-H, Singh DJ. Connecting thermoelectric performance and topological-insulator behavior: Bi2Te3 and Bi2Te2Se from first principles. Phys
Rev Appl 2015;3:014004.
[327] Sirikumara HI, Jayasekera T. Tunable indirect-direct transition of few-layer SnSe via interface engineering. J Phys-Condens Mater 2017;29:425501.
[328] Feng J, Qian X, Huang C-W, Li J. Strain-engineered artificial atom as a broad-spectrum solar energy funnel. Nat Photonics 2012;6:866–72.
[329] Akinwande D, Petrone N, Hone J. Two-dimensional flexible nanoelectronics. Nat Commun 2014;5:5678.
[330] Ziman JM. Electrons and phonons: the theory of transport phenomena in solids. Oxford University Press; 1960.
[331] Ju H, Kim M, Park D, Kim J. A strategy for low thermal conductivity and enhanced thermoelectric performance in SnSe: porous SnSe1–xSx nanosheets. Chem
Mater 2017;29:3228–36.
[332] Ju H, Kim J. Fabrication of conductive polymer/inorganic nanoparticles composite films: PEDOT:PSS with exfoliated tin selenide nanosheets for polymer-based
thermoelectric devices. Chem Eng J 2016;297:66–73.
[333] Liu S, Sun N, Sucharitakul S, Liu M, Gao X. IV-VI monochalcogenide SnSe nanostructures: synthesis, doping and thermoelectric properties. Bull Am Phys Soc
2017;62.
[334] Jiang J, Wong CPY, Zou J, Li S, Wang Q, Chen J, et al. Two-step fabrication of single-layer rectangular SnSe flakes. 2D Mater 2017;4:021026.
[335] Drozd VE, Nikiforova IO, Bogevolnov VB, Yafyasov AM, Filatova EO, Papazoglou D. ALD synthesis of SnSe layers and nanostructures. J Phys D Appl Phys
2009;42:125306.
[336] Subramanian B, Sanjeeviraja C, Jayachandran M. Brush plating of tin (II) selenide thin films. J Cryst Growth 2002;234:421–6.
[337] Zainal Z, Saravanan N, Anuar K, Hussein MZ, Yunus WMM. Chemical bath deposition of tin selenide thin films. Mater Sci Eng B-Adv 2004;107:181–5.
[338] Pejova B, Grozdanov I. Chemical synthesis, structural and optical properties of quantum sized semiconducting tin(II) selenide in thin film form. Thin Solid Films
2007;515:5203–11.
[339] Bindu K, Nair PK. Semiconducting tin selenide thin films prepared by heating Se–Sn layers. Semicond Sci Tech 2004;19:1348–53.
[340] Kumar N, Parihar U, Kumar R, Patel KJ, Panchal CJ, Padha N. Effect of film thickness on optical properties of tin selenide thin films prepared by thermal
evaporation for photovoltaic applications. Am J Mater Sci 2012;2:41–5.
[341] Kumar N, Sharma V, Padha N, Shah NM, Desai MS, Panchal CJ, et al. Influence of the substrate temperature on the structural, optical, and electrical properties
of tin selenide thin films deposited by thermal evaporation method. Cryst Res Technol 2010;45:53–8.
[342] Suguna P, Mangalaraj D, Narayandass SK, Meena P. Structure, composition, dielectric, and AC conduction studies on tin selenide films. Phys Status Solidi A
1996;155:405–16.

343
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[343] Kumar N, Sharma V, Parihar U, Sachdeva R, Padha N, Panchal C. Structure, optical and electrical characterization of tin selenide thin films deposited at room
temperature using thermal evaporation method. J Nano- Electron Phys 2011;3:117.
[344] Indirajith R, Srinivasan TP, Ramamurthi K, Gopalakrishnan R. Synthesis, deposition and characterization of tin selenide thin films by thermal evaporation
technique. Curr Appl Phys 2010;10:1402–6.
[345] John K, Pradeep B, Mathai E. Tin selenide (SnSe) thin films prepared by reactive evaporation. J Mater Sci 1994;29:1581–3.
[346] Jeong G, Kim J, Gunawan O, Pae SR, Kim SH, Song JY, et al. Preparation of single-phase SnSe thin-films and modification of electrical properties via
stoichiometry control for photovoltaic application. J Alloy Compd 2017;722:474–81.
[347] Zhang W, Yang Z, Liu J, Zhang L, Hui Z, Yu W, et al. Room temperature growth of nanocrystalline tin (II) selenide from aqueous solution. J Cryst Growth
2000;217:157–60.
[348] Zhang C, Yin H, Han M, Dai Z, Pang H, Zheng Y, et al. Two-dimensional tin selenide nanostructures for flexible all-solid-state supercapacitors. ACS Nano
2014;8:3761–70.
[349] Singh J, Bedi R. Thermally stimulated currents in epitaxially grown tin selenide films. Jpn J Appl Phys 1990;29:L869.
[350] Singh JP, Bedi RK. Tin selenide films grown by hot wall epitaxy. J Appl Phys 1990;68:2776.
[351] Singh J. Transport and optical properties of hot-wall-grown tin selenide films. J Mater Sci-Mater El 1991;2:105–8.
[352] Singh J, Bedi R. Electrical properties of flash-evaporated tin selenide films. Thin Solid Films 1991;199:9–12.
[353] Hema Chandra G, Naveen Kumar J, Madhusudhana Rao N, Uthanna S. Preparation and characterization of flash evaporated tin selenide thin films. J Cryst
Growth 2007;306:68–74.
[354] Pathinettam Padiyan D, Marikani A, Murali K. Electrical and photoelectrical properties of vacuum deposited SnSe thin films. Cryst Res Technol
2000;35:949–57.
[355] Rao TS, Chaudhuri A. Electrical and photoelectronic properties of SnSe thin films. J Phys D Appl Phys 1985;18:L35.
[356] Baxter C, McLennan W. Ovonic type switching in tin selenide thin films. J Vac Sci Technol 1975;12:110–3.
[357] Urmila K, Namitha T, Rajani J, Philip R, Pradeep B. Optoelectronic properties and Seebeck coefficient in SnSe thin films. J Semicond 2016;37:093002.
[358] Nair P, Martínez AK, García-Angelmo AR, Barrios-Salgado E, Nair M. Thermoelectric prospects of chemically deposited PbSe and SnSe thin films. Semicond Sci
Tech 2018;33:035004.
[359] Suen CH, Shi D, Su Y, Zhang Z, Chan CH, Tang X, et al. Enhanced thermoelectric properties of SnSe thin films grown by pulsed laser glancing-angle deposition. J
Materiomics 2017;3:293–8.
[360] Hiramatsu H, Ohta H, Hirano M, Hosono H. Heteroepitaxial growth of single-phase zinc blende ZnS films on transparent substrates by pulsed laser deposition
under H2S atmosphere. Solid State Commun 2002;124:411–5.
[361] Hiramatsu H, Ohta H, Suzuki T, Honjo C, Ikuhara Y, Ueda K, et al. Mechanism for heteroepitaxial growth of transparent p-type semiconductor: LaCuOS by
reactive solid-phase epitaxy. Cryst Growth Des 2004;4:301–7.
[362] Hubler GK. Pulsed laser deposition. Mrs Bull 1992;17:26–9.
[363] Horwitz JS. Pulsed-laser deposition. ASM International, Member/Customer Service Center, Materials Park, OH 44073–0002, USA, 1994 1994:621–6.
[364] Chrisey DB, Hubler GK. Pulsed laser deposition of thin films; 1994.
[365] Eason R. Pulsed laser deposition of thin films: applications-led growth of functional materials. John Wiley & Sons; 2007.
[366] Matthews BE, Holder AM, Schelhas LT, Siol S, May JW, Forkner MR, et al. Using heterostructural alloying to tune the structure and properties of the ther-
moelectric Sn1−xCaxSe. J Mater Chem A 2017;5:16873–82.
[367] Anwar S, Gowthamaraju S, Mishra BK, Singh SK, Shahid A. Spray pyrolysis deposited tin selenide thin films for thermoelectric applications. Mater Chem Phys
2015;153:236–42.
[368] Lukinskas A, Jasulaitienė V, Lukinskas P, Savickaja I, Kalinauskas P. Electrochemical formation of nanometric layers of tin selenides on Ti surface. Electrochim
Acta 2006;51:6171–8.
[369] Mathews NR. Electrodeposited tin selenide thin films for photovoltaic applications. Sol Energy 2012;86:1010–6.
[370] Engelken R, Berry A, Van Doren T, Boone J, Shahnazary A. Electrodeposition and analysis of tin selenide films. J Electrochem Soc 1986;133:581–5.
[371] Subramanian B, Mahalingam T, Sanjeeviraja C, Jayachandran M, Chockalingam MJ. Electrodeposition of Sn, Se, SnSe and the material properties of SnSe films.
Thin Solid Films 1999;357:119–24.
[372] Zainal Z, Ali AJ, Kassim A, Hussein MZ. Electrodeposition of tin selenide thin film semiconductor: effect of the electrolytes concentration on the film properties.
Sol Energ Mat Sol C 2003;79:125–32.
[373] Venkatasubramanian R, Siivola E, Colpitts T, O'quinn B. Thin-film thermoelectric devices with high room-temperature figures of merit. Nature
2001;413:597–602.
[374] Harman T, Cahn J, Logan M. Measurement of thermal conductivity by utilization of the Peltier effect. J Appl Phys 1959;30:1351–9.
[375] Harman T, Taylor P, Spears D, Walsh M. Thermoelectric quantum-dot superlattices with high ZT. J Electron Mater 2000;29:L1–2.
[376] Cahill DG. Analysis of heat flow in layered structures for time-domain thermoreflectance. Rev Sci Instrum 2004;75:5119–22.
[377] Chiritescu C, Cahill DG, Nguyen N, Johnson D, Bodapati A, Keblinski P, et al. Ultralow thermal conductivity in disordered, layered WSe2 crystals. Science
2007;315:351–3.
[378] Paddock CA, Eesley GL. Transient thermoreflectance from thin metal films. J Appl Phys 1986;60:285–90.
[379] Fernandes P, Sousa M, Salome PM, Leitao J, da Cunha A. Thermodynamic pathway for the formation of SnSe and SnSe2 polycrystalline thin films by sele-
nization of metal precursors. CrystEngComm 2013;15:10278–86.
[380] Herring C. Theory of the thermoelectric power of semiconductors. Phys Rev 1954;96:1163.
[381] Li Z, Bauers SR, Poudel N, Hamann D, Wang X, Choi DS, et al. Cross-plane seebeck coefficient measurement of misfit layered compounds (SnSe)n(TiSe2)n (n = 1,
3, 4, 5). Nano Lett 2017;17:1978–86.
[382] Martínez-Escobar D, Ramachandran M, Sánchez-Juárez A, Rios JSN. Optical and electrical properties of SnSe2 and SnSe thin films prepared by spray pyrolysis.
Thin Solid Films 2013;535:390–3.
[383] Yang S, Liu Y, Wu M, Zhao L-D, Lin Z, H-c Cheng, et al. Highly-anisotropic optical and electrical properties in layered SnSe. Nano Res 2018:554–64.
[384] Huang Y, Li L, Lin Y-H, Nan C-W. Liquid exfoliation few-layer SnSe nanosheets with tunable band gap. J Phys Chem C 2017;121:17530–7.
[385] Li L, Chen Z, Hu Y, Wang X, Zhang T, Chen W, et al. Single-layer single-crystalline SnSe nanosheets. J Am Chem Soc 2013;135:1213–6.
[386] Luo S, Qi X, Yao H, Ren X, Chen Q, Zhong J. Temperature-dependent raman responses of the vapor-deposited tin selenide ultrathin flakes. J Phys Chem C
2017;121:4674–9.
[387] Xu X, Song Q, Wang H, Li P, Zhang K, Wang Y, et al. In-plane anisotropies of polarized raman response and electrical conductivity in layered tin selenide. ACS
Appl Mater Interf 2017;9:12601–7.
[388] Pei T, Bao L, Ma R, Song S, Ge B, Wu L, et al. Epitaxy of ultrathin SnSe single crystals on polydimethylsiloxane: in-plane electrical anisotropy and gate-tunable
thermopower. Adv Electron Mater 2016;2:1600292.
[389] Zhou X, Gan L, Tian W, Zhang Q, Jin S, Li H, et al. Ultrathin SnSe2 flakes grown by chemical vapor deposition for high-performance photodetectors. Adv Mater
2015;27:8035–41.
[390] Wang Z, Xu K, Li Y, Zhan X, Safdar M, Wang Q, et al. Role of Ga vacancy on a multilayer GaTe phototransistor. ACS Nano 2014;8:4859–65.
[391] Ju H, Kim J. Chemically exfoliated SnSe nanosheets and their SnSe/poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) composite films for polymer
based thermoelectric applications. ACS Nano 2016;10:5730–9.
[392] Ko J, Kim J-Y, Choi S-M, Lim YS, Seo W-S, Lee KH. Nanograined thermoelectric Bi2Te2.7Se0.3 with ultralow phonon transport prepared from chemically
exfoliated nanoplatelets. J Mater Chem A 2013;1:12791–6.
[393] Ren L, Qi X, Liu Y, Hao G, Huang Z, Zou X, et al. Large-scale production of ultrathin topological insulator bismuth telluride nanosheets by a hydrothermal
intercalation and exfoliation route. J Mater Chem 2012;22:4921–6.

344
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

[394] Zeng Z, Yin Z, Huang X, Li H, He Q, Lu G, et al. Single-layer semiconducting nanosheets: high-yield preparation and device fabrication. Angew Chem Int Ed
2011;50:11093–7.
[395] Ju H, Kim K, Park D, Kim J. Fabrication of porous SnSeS nanosheets with controlled porosity and their enhanced thermoelectric performance. Chem Eng J
2018;335:560–6.
[396] Zhang Q, Wang W, Li J, Zhu J, Wang L, Zhu M, et al. Preparation and thermoelectric properties of multi-walled carbon nanotube/polyaniline hybrid nano-
composites. J Mater Chem A 2013;1:12109–14.
[397] Wang J, Cai K, Shen S, Yin J. Preparation and thermoelectric properties of multi-walled carbon nanotubes/polypyrrole composites. Synthetic Met
2014;195:132–6.
[398] Yue R, Xu J. Poly (3, 4-ethylenedioxythiophene) as promising organic thermoelectric materials: a mini-review. Synthetic Met 2012;162:912–7.
[399] Mateeva N, Niculescu H, Schlenoff J, Testardi L. Correlation of Seebeck coefficient and electric conductivity in polyaniline and polypyrrole. J Appl Phys
1998;83:3111–7.
[400] Yan H, Sada N, Toshima N. Thermal transporting properties of electrically conductive polyaniline films as organic thermoelectric materials. J Therm Anal
Calorim 2002;69:881–7.
[401] Kim YH, Sachse C, Machala ML, May C, Müller-Meskamp L, Leo K. Highly conductive PEDOT: PSS electrode with optimized solvent and thermal post-treatment
for ITO-free organic solar cells. Adv Funct Mater 2011;21:1076–81.
[402] Luo J, Billep D, Waechtler T, Otto T, Toader M, Gordan O, et al. Enhancement of the thermoelectric properties of PEDOT: PSS thin films by post-treatment. J
Mater Chem A 2013;1:7576–83.
[403] Kim G, Shao L, Zhang K, Pipe KP. Engineered doping of organic semiconductors for enhanced thermoelectric efficiency. Nat Mater 2013;12:719.
[404] Na S-I, Wang G, Kim S-S, Kim T-W, Oh S-H, Yu B-K, et al. Evolution of nanomorphology and anisotropic conductivity in solvent-modified PEDOT: PSS films for
polymeric anodes of polymer solar cells. J Mater Chem 2009;19:9045–53.
[405] He M, Qiu F, Lin Z. Towards high-performance polymer-based thermoelectric materials. Energy Environ Sci 2013;6:1352–61.
[406] He M, Ge J, Lin Z, Feng X, Wang X, Lu H, et al. Thermopower enhancement in conducting polymer nanocomposites via carrier energy scattering at the
organic–inorganic semiconductor interface. Energy Environ Sci 2012;5:8351–8.
[407] Zhang B, Sun J, Katz H, Fang F, Opila R. Promising thermoelectric properties of commercial PEDOT: PSS materials and their Bi2Te3 powder composites. ACS
Appl Mater Interf 2010;2:3170–8.
[408] Wang Y, Cai K, Yao X. Facile fabrication and thermoelectric properties of PbTe-modified poly (3, 4-ethylenedioxythiophene) nanotubes. ACS Appl Mater Interf
2011;3:1163–6.
[409] Ding Z, Viculis L, Nakawatase J, Kaner R. Intercalation and solution processing of bismuth telluride and bismuth selenide. Adv Mater 2001;13:797–800.
[410] Jiang Q, Liu C, Song H, Shi H, Yao Y, Xu J, et al. Improved thermoelectric performance of PEDOT: PSS films prepared by polar-solvent vapor annealing method.
J Mater Sci-Mater El 2013;24:4240–6.
[411] Tsai TC, Chen TH, Chang HC, Chen CH, Huang YC, Whang WT. A facile surface treatment utilizing binary mixtures of ammonium salts and polar solvents for
multiply enhancing thermoelectric PEDOT: PSS films. J Polym Sci Part A: Polym Chem 2014;52:3303–6.
[412] Wang M, Zhou M, Zhu L, Li Q, Jiang C. Enhanced polymer solar cells efficiency by surface coating of the PEDOT: PSS with polar solvent. Sol Energy
2016;129:175–83.
[413] Coates NE, Yee SK, McCulloch B, See KC, Majumdar A, Segalman RA, et al. Effect of interfacial properties on polymer–nanocrystal thermoelectric transport. Adv
Mater 2013;25:1629–33.
[414] Gelbstein Y. Thermoelectric power and structural properties in two-phase Sn/SnTe alloys. J Appl Phys 2009;105:023713.
[415] Wei Q, Mukaida M, Naitoh Y, Ishida T. Morphological change and mobility enhancement in PEDOT: PSS by adding co-solvents. Adv Mater 2013;25:2831–6.
[416] Julien C, Eddrief M, Samaras I, Balkanski M. Optical and electrical characterizations of SnSe, SnS2 and SnSe2 single crystals. Mater Sci Eng B-Adv 1992;15:70–2.
[417] Elkorashy A. Optical constants of tin selenide single crystals by interference method. J Phys Chem Solids 1990;51:289–96.
[418] Elkorashy A. Optical absorption in tin monoselenide single crystal. J Phys Chem Solids 1986;47:497–500.
[419] Jin M, Shao H, Hu H, Li D, Shen H, Xu J, et al. Growth and characterization of large size undoped p-type SnSe single crystal by Horizontal Bridgman method. J
Alloy Compd 2017;712:857–62.
[420] Nassary M. The electrical conduction mechanisms and thermoelectric power of SnSe single crystals. Turk J Phys 2009;33:201–8.
[421] Maier H, Daniel D. SnSe single crystals: sublimation growth, deviation from stoichiometry and electrical properties. J Electron Mater 1977;6:693–704.
[422] Serrano-Sánchez F, Nemes N, Dura O, Fernandez-Diaz M, Martínez J, Alonso J. Structural phase transition in polycrystalline SnSe: a neutron diffraction study in
correlation with thermoelectric properties. J Appl Crystallogr 2016;49:2138–44.
[423] Chen W-H, Yang Z-R, Lin F-H, Liu C-J. Nanostructured SnSe: hydrothermal synthesis and disorder-induced enhancement of thermoelectric properties at
medium temperatures. J Mater Sci 2017;52:9728–38.
[424] Liu H, Zhang X, Li S, Zhou Z, Liu Y, Zhang J. Synthesis and thermoelectric properties of SnSe by mechanical alloying and spark plasma sintering method. J
Electron Mater 2017;46:2629–33.
[425] Peng K, Wu H, Yan Y, Guo L, Wang G, Lu X, et al. Grain size optimization for high-performance polycrystalline SnSe thermoelectrics. J Mater Chem A
2017;5:14053–60.
[426] Yang SD, Nutor RK, Chen ZJ, Zheng H, Wu HF, Si JX. Influence of sodium chloride doping on thermoelectric properties of p-type SnSe. J Electron Mater
2017;46:6662–8.
[427] Lee ST, Kim MJ, Lee G-G, Kim SG, Lee S, Seo W-S, et al. Effects of Sn-deficiency on thermoelectric properties of polycrystalline Sn1-xSe compounds. Curr Appl
Phys 2017;17:732–7.
[428] Yang K, Ren J-C, Qiu H, Wang J-S. Phonon-driven electron scattering and magnetothermoelectric effect in two-dimensional tin selenide. J Phys-Condens Mater
2018;30:055301.
[429] Parenteau M, Carlone C. Influence of temperature and pressure on the electronic transitions in SnS and SnSe semiconductors. Phys Rev B 1990;41:5227–34.

Assoc. Prof. Zhi-Gang Chen is currently an Associate Professor of Energy Materials at the University of Southern Queensland. He
received his Ph.D. in Materials Science and Engineering from the Institute of Metal Research, Chinese Academy of Science, in 2008. After
his Ph.D., he worked at the University of Queensland for seven years with various prestigious fellowships, including ARC Postdoctoral
Fellowship, QLD Smart Future Fellowship. In 2012, he won a Queensland International Fellowship to undertake a collaborative research
at California Institute of Technology. His research concentrates on smart functional nanomaterials for thermoelectrics and nanoelec-
tronics, from synthesizing materials to understanding their underlying physics and chemistry.

345
Z.-G. Chen et al. Progress in Materials Science 97 (2018) 283–346

Xiaolei Shi received his Bachelor’s degree from the University of Science and Technology in Beijing, China, in 2008, and Master of
Philosophy in the field of applications on atomic force microscopy from the University of Science and Technology in Beijing, China, in
2011. After his MPhi, he worked as an engineer at Tsinghua University in the field of chemical and mechanical planarization. In 2015, he
won an International Postgraduate Research Scholarship to undertake his Ph.D project under the supervision of Prof. Jin Zou and Assoc.
Prof. Zhi-Gang Chen at the University of Queensland with a research focus on the development of high-performance stannous selenide
thermoelectrics and underlying mechanisms.

Prof. Li -Dong Zhao is a full Professor of Materials Science and Engineering at Beihang University, China. He received his B.E. and M.E.
degrees in materials science from Liaoning Technical University and his Ph.D. degree in materials science from the University of Science
and Technology Beijing, China, in 2009. He was a postdoctoral research fellow in the ICMMO at the University of Paris ‐ Sud from 2009
to 2011 and continued as a postdoctoral research fellow in the Department of Chemistry at Northwestern University from 2011 to 2014.
His research interests include the fabrication and characterization of layered structural thermoelectrics, superconductors, and thermal
barrier coatings.

Prof. Jin Zou is currently the Chair of Nanoscience at the University of Queensland. He received his Ph.D. in Materials Physics in late
1993 from Sydney University, Australia, and worked there for 10 years with various prestigious fellowships, including an Australian
Government’s Queen Elizabeth II Fellowship. In the second half of 2003, Professor Zou moved to the University of Queensland and
continued his research in the field of semiconductor nanostructures for energy-related applications.

346

Potrebbero piacerti anche