ArticlePDF Available

Abstract and Figures

An aqueous solution method is developed for the facile synthesis of Cl-containing SnSe nanoparticles in 10 g quantities per batch. The particle size and Cl concentration of the nanoparticles can be efficiently tuned as a function of reaction duration. Hot pressing produces n-type Cl-doped SnSe nanostructured compacts with thermoelectric power factors optimized via control of Cl dopant concentration. This approach, combining an energy-efficient solution synthesis with hot pressing, provides a simple, rapid, and low-cost route to high performance n-type SnSe thermoelectric materials.
Content may be subject to copyright.
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (1 of 7) 1602328
wileyonlinelibrary.com
Chlorine-Enabled Electron Doping in Solution-Synthesized
SnSe Thermoelectric Nanomaterials
Guang Han, Srinivas R. Popuri, Heather F. Greer, Lourdes F. Llin, Jan-Willem G. Bos,
Wuzong Zhou, Douglas J. Paul, Hervé Ménard, Andrew R. Knox, Andrea Montecucco,
Jonathan Siviter, Elena A. Man, Wen-guang Li, Manosh C. Paul, Min Gao, Tracy Sweet,
Robert Freer, Feridoon Azough, Hasan Baig, Tapas K. Mallick, and Duncan H. Gregory*
DOI: 10.1002/aenm.201602328
precursors,[4] doping proves challenging
for solution-synthesized MC nanostruc-
tures.[5] Recently, post-synthesis halide
treatment of nanocrystals in solution has
been developed which involves switching
halogens for long chain surfactant mol-
ecules absorbed on the surface.[2,6] In fact,
sorption of halogens can be realized as
part of a one-pot synthesis using metal
halide precursors.[7] Although this strategy
was initially developed for passivation of
MC quantum dots against oxidation,[2,6,7]
annealing or hot pressing halogen-coated
nanoparticles allows halides to diffuse into the MC lattice and
substitute for chalcogenide anions.[8] However, controlling
doping levels is not straightforward and such methods can
introduce rather high halide concentrations in small nanocrys-
tals leading to reduced electrical conductivity.[8b] Hence exerting
control over dopant concentration without sacrificing electrical
performance is imperative.
Thermoelectrics realize direct interconversion between
thermal and electric energy, thus providing an important route
An aqueous solution method is developed for the facile synthesis of Cl-con-
taining SnSe nanoparticles in 10 g quantities per batch. The particle size and
Cl concentration of the nanoparticles can be efficiently tuned as a function of
reaction duration. Hot pressing produces n-type Cl-doped SnSe nanostruc-
tured compacts with thermoelectric power factors optimized via control of Cl
dopant concentration. This approach, combining an energy-efficient solution
synthesis with hot pressing, provides a simple, rapid, and low-cost route to
high performance n-type SnSe thermoelectric materials.
Doping plays a vital role in modifying the electronic properties
of semiconductors and is pivotal for (opto)electronics,[1] photo-
voltaics (PV),[2] and thermoelectrics.[3] Metal chalcogenides
(MCs) form a diversity of functional materials well-suited to
such applications. Halogen doping in MCs has proven effec-
tive to realize n-type conducting behavior and tune carrier
concentrations.[2–4] Enhanced thermoelectric and PV perfor-
mance can result.[2–4] While halogens can be readily doped into
bulk MCs by high-temperature synthesis using metal halide
Dr. G. Han, Prof. D. H. Gregory
WestCHEM
School of Chemistry
University of Glasgow
Glasgow G12 8QQ, UK
E-mail: Duncan.Gregory@glasgow.ac.uk
Dr. S. R. Popuri, Dr. J.-W. G. Bos
Institute of Chemical Sciences and Centre for Advanced Energy
Storage and Recovery
School of Engineering and Physical Sciences
Heriot-Watt University
Edinburgh EH14 4AS, UK
Dr. H. F. Greer, Prof. W. Z. Zhou
EaStCHEM
School of Chemistry
University of St Andrews
St Andrews, Fife KY16 9ST, UK
Dr. L. F. Llin, Prof. D. J. Paul, Prof. A. R Knox, Dr. A. Montecucco,
Dr. J. Siviter, Dr. E. A. Man, Dr. W.-g. Li, Dr. M. C. Paul
School of Engineering
University of Glasgow
Glasgow G12 8QQ, UK
Dr. H. Ménard
Sasol (UK) Ltd
St Andrews, Fife KY16 9ST, UK
Prof. M. Gao, Dr. T. Sweet
School of Engineering
Cardiff University
Cardiff CF24 3AA, UK
Prof. R. Freer, Dr. F. Azough
School of Materials
University of Manchester
Manchester M13 9PL, UK
Dr. H. Baig, Prof. T. K. Mallick
Environment and Sustainability Institute
University of Exeter
Penryn Campus, Penryn TR10 9FE, UK
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.dewww.advancedsciencenews.com
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1602328 (2 of 7) wileyonlinelibrary.com
to produce useful electricity from waste heat and to perform
refrigeration (via the Seebeck and Peltier effects, respectively).[9]
SnSe is a layer-structured MC semiconductor and potentially
useful thermoelectric material given its excel-
lent energy conversion efficiency, relatively
low toxicity, and the high earth-abundance of
the component elements.[10] Most research to
date has concentrated on p-type SnSe.[10,11]
Conversely n-type SnSe is difficult to achieve;
only I and BiCl3 have been used successfully
to dope bulk SnSe with electrons. Moreover,
high-temperature, energy-intensive processes
are needed to achieve this.[12] There are no
reports of solution-synthesized SnSe nano-
structures with tunable n-type conducting
behavior. Before the potential of SnSe can be
fully realized, it is critical to develop a cost-
effective and large-scale synthesis of high
performing n-type SnSe, to complement
existing p-type materials.
In this study, we demonstrate the intro-
duction of Cl to SnSe nanoparticles by a
one-pot in situ solution approach to prepare
>10 g of doped SnSe nanoparticles on short
timescales (Scheme 1; Figure S1, Supporting
Information). The strategy exploits the nucle-
ophilic nature of the halide anion and the
electrophilicity of coordinatively unsaturated
metal cations at the nanoparticle surface,[6]
coupled with the acidic conditions that pro-
mote the formation of metal–halide bonds by
ligand replacement.[8a] The simple solution
synthesis is achieved using aqueous SnCl2
both as reactant and Cl source and citric
acid both as surfactant to restrict particle
growth and means to control pH. Controlling
the reaction duration allows us to engineer
nanoparticle size and regulate the Cl dopant
level. The nanoparticles can be hot-pressed
into Cl-doped SnSe dense pellets with controllable dopant con-
centration and consistent n-type conducting behavior.
Injection of an NaHSe aqueous solution into an SnCl2 solu-
tion (26:1 molar ratio of citric acid:SnCl2) leads to the imme-
diate formation of an SnSe precipitate (Equation (1))
+→++Na
HSeSnClSnSeNaClHCl
2 (1)
Boiling the suspension for 2 h generates crystalline, phase-
pure SnSe nanoparticles. Powder X-ray diffraction (PXD) pat-
terns can be exclusively indexed to orthorhombic SnSe (ICDD
card No. 48–1224).[13] Rietveld refinement against PXD data
(Figure 1a; Tables S1 and S2, Supporting Information) confirms
that the single phase SnSe product crystallizes in orthorhombic
space group Pnma, with a = 11.5424(8) Å, b = 4.1775(4) Å, and
c = 4.3841(5) Å. Scanning electron microscopy (SEM) images
(Figure S2a,b, Supporting Information) reveal that the product
is an agglomeration of nanoparticles with individual sizes of
15–55 nm. Energy dispersive X-ray spectroscopy (EDS) spectra
(Figure S2c, Supporting Information) taken across the samples
as point and area scans consistently generate Sn:Se:Cl atomic
ratios of 50.6(5):48.8(5):0.6(1). The existence of Cl should result
from the interaction of nucleophilic Cl and electrophilic Sn2+
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.de www.advancedsciencenews.com
Scheme 1. The strategy to fabricate n-type SnSe nanostructured pellets
utilizes Cl concentration and nanoparticle size from solution-synthesis.
Figure 1. Characterization of SnSe nanoparticles synthesized after 2 h. a) Profile plot from Riet-
veld refinement against PXD data. b) TEM image and corresponding SAED pattern collected
from the nanoparticles. c) HRTEM image of an individual nanoparticle with the (201) d-spacing
indicated. d) EDS mapping of a nanoparticle cluster. The four panels in (d) are (clockwise from
top left) high angle annular dark field (HAADF) image, elemental mapping for Sn (red), Cl
(yellow), and Se (green).
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (3 of 7) 1602328
wileyonlinelibrary.com
at the nanoparticle surface[6] together with the replacement
of ligated citric acid by Cl in the acidic environment during
solution heating.[8a] Transmission electron microscopy (TEM)
images (Figure 1b; Figure S2d, Supporting Information) con-
firm that SnSe nanoparticles assemble into clusters with an
average individual particle size of 35 nm. Selected area electron
diffraction (SAED) patterns (Figure 1b) collected from a section
of such a cluster reveal the polycrystalline nature of the SnSe
particles. High-resolution TEM (HRTEM) images (Figure 1c)
nevertheless demonstrate the single crystalline nature of indi-
vidual nanoparticles. Elemental mapping (Figure 1d) confirms
the existence and uniform distribution of Sn, Se, and Cl in the
nanoparticles.
Controlling the synthesis duration can predetermine both
the particle size and Cl content of the SnSe nanoparticles. To
illustrate this, we synthesized materials from 1 min, 5 min, and
24 h of solution heating. PXD (Figures S3 and S4; Table S1,
Supporting Information) reveals each product to be single-
phase with the orthorhombic SnSe structure. The refined cell
volumes increase slightly with reaction time and the Bragg half-
widths decrease gradually as the reaction duration increases,
indicating likely crystallite growth. SEM (Figure S5a–f, Sup-
porting Information) and TEM (Figure 2a,b: Figure S6a,
Supporting Information) images show that the products are
composed of nanoparticles, and the average particle size
increases from 25 nm through 30 to 50 nm, as heating is
extended. EDS (Figure S5g–i, Supporting
Information) confirms the existence of Cl
in all the samples and shows increased
Cl levels for shorter reaction times; spe-
cifically, the Sn:Se:Cl atomic ratios are
48.2(5):50.5(5):1.3(2), 51.6(5):47.7(5):0.7(1),
and 51.5(5):48.2(5):0.3(1) for nanoparticles
synthesized after 1 min, 5 min, and 24 h,
respectively. SAED patterns (Figure 2a,b;
Figure S6b, Supporting Information) con-
firm the polycrystalline nature of the SnSe
nanoparticles. Furthermore, HRTEM
(Figure 2c; Figure S6c,d, Supporting Infor-
mation) reveals that the products synthe-
sized after 1 and 5 min have relatively poor
crystallinity, and nanoparticles with sizes of
2–4 nm are attached on the surface of larger
particles. When the reaction time is increased
from 2 to 24 h (Figure 2d), individual parti-
cles become single crystalline. This suggests
that the aggregation and coalescence of small
nanoparticles leads to the formation of larger
single crystalline nanoparticles.[14]
The ability to prepare SnSe nanomate-
rials in >10 g quantities allowed the facile
fabrication of SnSe pellets via hot pressing
without the variations in sample morphology
that could ensue from repeated sample
preparation. Pellets with 95% of the SnSe
theoretical density, consolidated from 2 h
solution-synthesized powder, were obtained
(denoted 1). Rietveld refinement (Figure S7;
Table S3, Supporting Information) shows
that the pellets are composed principally of orthorhombic SnSe
(79.1(1) wt%) but also of two minority phases of trigonal SnSe2
(11.1(3) wt%) and tetragonal SnO2 (9.8(2) wt%). This indicates
that a small proportion of SnSe was oxidized to SnO2 and SnSe2
during the hot pressing process (2SnSe + O2 SnO2 + SnSe2).
A series of subsequent experiments (see Figures S8–S15;
Tables S4–S6, Supporting Information) confirmed that the oxi-
dation was less in nanomaterials prepared at longer reaction
durations. Both particle size (and by inference surface area) and
the relative amount of surface carboxyl groups could be traced
as contributors to the oxidation process.
Preferred orientation of both SnSe and SnSe2 crystallites is
evidenced by the increased intensity of the (h00) and (00l) reflec-
tions respectively in PXD patterns from the pellets (Figure S16,
Supporting Information). This suggests that the longest crystal-
lographic axes of the respective cells align parallel to the hot
pressing direction. Considering that the magnitude of the elec-
trical conductivity is highest perpendicular to these long axes
in both SnSe and SnSe2,[10a,15] such an orientation should be
beneficial in enhancing electrical conductivity perpendicular to
the hot pressing direction of the pellets.
SEM and TEM images (Figure S17a,b and S18a, Supporting
Information) show that 1 is composed of densely packed plates
with almost uniformly distributed particles. EDS (Figure S17c–g,
Supporting Information) confirms the existence and uniform
distribution of Cl in 1. SAED patterns (Figure S18b, Supporting
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.dewww.advancedsciencenews.com
Figure 2. TEM characterization of SnSe nanoparticles synthesized after a,c) 1 min and b,d) 24 h.
a,b) TEM images and corresponding SAED patterns collected from the particles. c,d) HRTEM
images of individual nanoparticles with selected d-spacings indicated.
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1602328 (4 of 7) wileyonlinelibrary.com
Information) taken from a number of plates and nanoparticles
from 1 confirm the presence of the above-mentioned three
phases: SnSe, SnSe2, and SnO2. HRTEM images demonstrate
that the predominant plate-like structures in 1 are crystalline
SnSe (Figure S18c, Supporting Information). High-resolution
images also show that some of the smaller irregular nano-
plates are formed by SnSe2 (Figure S18d, Supporting Informa-
tion) while some nano particles of SnO2 are distributed among
the SnSe plates (Figure S18e, Supporting Information). Ther-
mogravimetric-differential thermal analysis of 1 under argon
(Figure S19, Supporting Information) shows negligible weight
loss below 500 °C, but reveals that decomposition begins above
this temperature and corresponds to an endothermic Se subli-
mation process.
For comparison, the nanomaterials prepared over 1 min,
5 min, and 24 h were also hot pressed into pellets (denoted
2, 3, 4, respectively) using the same processing parame-
ters, achieving 85%, 90%, and 92% of the SnSe theoretical
density, respectively. Rietveld refinement against PXD data
(Figure S20; Tables S7–S9, Supporting Information) shows
that the SnSe phase fraction increases from 71 through 72 to
90 wt% for 2, 3, and 4, respectively, again indicating a direct
correlation between synthesis time (and particle size/surface
citric acid amount) and the tendency to oxidation. HRTEM
on 3 confirms that SnO2 nanoparticles are distributed in close
proximity to the SnSe plates (Figure S21, Supporting Infor-
mation). EDS (Figure S22, Supporting Information) confirms
that the pellets contain Cl at levels consistent with the corre-
sponding solution synthesized SnSe nanoparticles, suggesting
no loss during the hot pressing process. X-ray photoelectron
spectroscopy (XPS) was used to verify the presence of chlorine
and analysis of 3 (Figure S23, Supporting Information) shows
that the peaks at 200.5 and 198.9 eV can be assigned to Cl
2p1/2 and Cl 2p3/2 states, respectively, indicating that Cl exists
in the form of Cl.[7a,16] This implies electron doping indeed
originates from the halide on substitution for Se2.[16] The
indirect optical bandgap from diffuse reflectance (DR) UV–Vis
spectra (Figure S24, Supporting Information) narrows from
0.8 through 0.75 to 0.7 eV, when the Cl concentration is
increased from 0.3% (4) through 0.7%–0.6% (3, 1) to 1.3%
(2), respectively. This insinuates that the indirect bandgap
of SnSe is reduced slightly but significantly by increased Cl
doping. A similar bandgap narrowing was observed in I-doped
SnSe and is expected to improve the electrical conductivity in
SnSe.[12a]
We selected 1, 3, and 4 for electrical measurements due to the
relatively low percentage of SnO2 and SnSe2 components and
the high density achieved (90%). Hall measurements (Table 1)
give a clear correlation between the Cl and carrier concentra-
tions (where the majority carriers are electrons). 1 and 3 have
higher carrier concentrations than 4 due to the twofold increase
in Cl content, while 3 has a lower carrier concentration than
1, which is probably related to the higher level of impurities.
The contrast in the temperature-dependent Seebeck coefficient
(S) for the different pellets is striking (Figure 3a; Table 1). The
variation in the absolute value of S with temperature for 1 and
3 is very similar increasing from 300 K to reach a maximum at
410–425 K before decreasing by 540 K. The decreasing value
of S at higher temperature could be due to thermal excitation
of minority carriers (holes) that are related to intrinsic defects
in SnSe (e.g., Sn vacancies)[17] as manifested by the significantly
enhanced electrical conductivity (Figure 3b). The absolute value
of 1 is slightly lower than that of 3 at 300 K, but S for both 1 and
3 are negative within the whole temperature range, indicating
n-type behavior consistent with Cl doping. By comparison, 4,
with the lowest Cl doping level, shows n-type behavior at 300 K
and transforms to p-type behavior at 475 K with S 75 µV K1
at 540 K. The n- to p-type transition could also be related to the
thermal excitation of holes at high temperature. We also note
that the impurity phases, SnSe2 and SnO2, are both intrinsic
n-type semiconductors.[16,18]
The electrical conductivity (
σ
) of 1 (Figure 3b) increases from
255 S m1 at 300 K to 910 S m1 at 540 K. With a similar
Cl doping concentration and indirect bandgap to 1, the
σ
of 3
is slightly lower (from 185 S m1 at 300 K to 685 S m1 at
540 K), probably due to the increase in the less conductive SnSe2
and SnO2 components,[16,18b] together with the increased car-
rier scattering from SnO2 nanoparticles. This is consistent with
the higher carrier concentration and mobility in 1 (Table 1). By
contrast, with the lowest Cl doping level, 4 exhibits the lowest
carrier concentration and
σ
among the three pellets, although
it is the least oxidized. In fact, 1 and 3 demonstrate higher
σ
values than bulk I-doped SnSe materials with similar doping
concentrations within the same temperature range (e.g.,
σ
for
SnSe0.98I0.02 increased from 0.23 S m1 at 300 K to 105 S m1
at 565 K) and comparable
σ
to Bi and Cl codoped materials
(e.g., SnSe0.95-0.2 mol% BiCl3 yielded
σ
values of 1170 S m1
at 300 K and 1815 S m1 at 560 K).[12]
In an attempt to understand the possible origins of the
n-type conducting behavior more fully, we also measured the
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.de www.advancedsciencenews.com
Table 1. A summary of Cl concentration, phase fraction, electrical properties, n–p transformation temperature, room temperature Hall carrier concen-
tration (nH), and mobility (
µ
H) of the pellets.
Pellet at% Cl S300 K
[µV K1]
σ
300 K
[S m1]
Tn–p
[K]
wt% SnO2wt% SnSe2nH
[1018 cm3]
µ
H
[cm2 V1 s1]
1 0.6(1) 265 255 a) 9.8(2) 11.1(3) 6.43 3.60
3 0.7(1) 295 185 a) 13.0(2) 15.3(3) 3.47 2.66
4 0.3(1) 145 55 475 3.5(2) 6.1(2) 2.56 0.85
SF1 0.3(1) 175b) 55b) 520 7.9(1) 6.7(2) 1.68 2.14
SF2 0.1(1) 95b) 50b) 400 5.5(1) 6.8(2) 1.24 1.62
a)No transition below 550 K; b)Values obtained at 325 K.
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (5 of 7) 1602328
wileyonlinelibrary.com
thermoelectric performance of two pellets synthesized using
hydrochloric acid in place of citric acid (SF1 and SF2; see Sup-
porting Information) for comparison (Figure 3; Table 1). EDS
shows that these surfactant-free samples have larger particle size
and lower Cl concentrations compared to their citric acid “ana-
logues,” 1 and 4. (The Sn:Se:Cl ratios are 51.7(5):48.0(5):0.3(1)
and 51.5(5):48.4(5):0.1(1) for SF1 and SF2, respectively;
Figure S14, Supporting Information). Three notable compari-
sons can be made: (1)SF2 contains slightly more SnO2 and
SnSe2 than 4 but contains less dopant Cl. SF2 has a lower room
temperature carrier concentration and
σ
, slightly lower magni-
tude Seebeck coefficient and transforms from n- to p-type at a
lower temperature than 4; (2) SF1 has a similar Cl concentration
but notably contains more SnO2 than 4. SF1 transforms from
n- to p-type at a higher temperature than 4; (3) 1 and 3 remain
n-type below 550 K with similar Seebeck coefficients, although
the electrical conductivity of 1 is higher than 3. Although there
are likely to be other contributing factors, the results indicate
that as Cl doping levels increase so does the electrical conduc-
tivity and the temperature of the p–n transition, both obser-
vations being consistent with a higher number of negative
charge carriers. Nevertheless, the presence of SnO2 (and SnSe2)
clearly also has an effect on the electrical properties, apparently
reducing the conductivity and increasing the temperature of the
n–p transition. These observations would certainly be consistent
with the presence of SnO2 as a wide band gap, n-type semicon-
ductor (Eg = 3.6 eV; typically
σ
4 S m1, S 200 µV K1 at
300 K, depending on oxygen vacancy concentration).[18] (SnSe2
has a gap of 1.6 eV,[19]
σ
of 170 S m1,[16] S 238 µV K1 at
300 K.[20]) Moreover, controlled doping of Cl is clearly very
effective in producing high performance n-type SnSe, but oxide
impurities need to be minimized to optimize this performance.
It is also worth noting that the absence of surfactant in the prep-
aration of SF1 and SF2 ultimately leads to significant improve-
ments in
σ
, especially at higher temperature (cf. 4).[11d]
The combination of better
σ
values coupled with high values
of S leads to higher power factors (S2
σ
) in 1 (0.018 mW m1 K2
at 300 K to 0.068 mW m1 K2 at 530 K) (Figure 3c). 3 shows
slightly lower values than 1 (S2
σ
0.016 mW m1 K2 at 300 K
and 0.054 mW m1 K2 at 525 K) as noted above. In contrast,
the S2
σ
values for 4 are much lower (0.001 mW m1 K2 at
300 K and reaching only 0.002 mW m1 K2 at 540 K). SF1
and SF2 achieve similar S2
σ
values at room temperature where
they are both n-type, whereas SF2 has a higher power factor at
525 K (where it is p-type). The contrast in performance between
samples underscores the importance of being able to tune the
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.dewww.advancedsciencenews.com
Figure 3. Electrical properties of SnSe pellets 1, 3, 4, SF1, and SF2 measured perpendicular to the hot pressing direction: a) the Seebeck coefficient,
b) the electrical conductivity, and c) the power factor as a function of temperature.
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1602328 (6 of 7) wileyonlinelibrary.com Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.de www.advancedsciencenews.com
degree of Cl doping and to control the pellet phase composi-
tion during fabrication. It is especially notable that the power
factor for 1 compares very favorably with those for I-doped
SnSe (e.g., SnSe0.98I0.02, with S2
σ
of 0.016 mW m1 K2 at
565 K) and co-doped SnSe (e.g., SnSe0.95-0.2 mol% BiCl3, with
an S2
σ
of 0.104 mW m1 K2 at 515 K) bulk materials with
similar doping levels within the same temperature range.[12]
If oxidation could be reduced, it might conceivable to surpass
such values. Hence, it is achievable to produce high performing
n-type SnSe materials in bulk quantities via energy-efficient,
sustainable methods (Figure S25, Supporting Information). In
principle, it should be possible to produce new co-doped SnSe
nanomaterials controllably (e.g., with both Bi- and Cl-dopants
among others) with only minor adaptations to the present syn-
thesis method.
Preliminary thermal conductivity measurements (
κ
) per-
formed on 1 and 4 along the direction parallel to pressing
(Figure S26, Supporting Information) are also encouraging.
κ
for 1 (4) decreases from 0.89 (0.72) W m1 K1 at 300 K
to 0.62 (0.40) W m1 K1 at 540 K. The higher
κ
for 1 could
be due to its higher percentage of more thermally conductive
SnO2. However,
κ
values for both 1 and 4 are still relatively low
compared to other examples of n-type polycrystalline SnSe and
to p-type single crystals (Figure S26d,e, Supporting Informa-
tion). This could be due to enhanced phonon scattering either
from the SnO2 nanoinclusions in these materials or as a result
of the nanostructuring of SnSe itself.
In summary, a simple, quick, low-cost solution synthesis
produces Cl-containing SnSe nanoparticles in gram quantities
(>10 g per run for a 2 h growth). Such nanoparticles have been
consolidated into n-type Cl-doped SnSe nanostructured pel-
lets, whose thermoelectric power factors can be significantly
improved by optimizing the Cl doping level. This study not
only provides a convenient method for the large-scale synthesis
of SnSe nanostructures, but also demonstrates a facile and reli-
able route to engineer n-type SnSe with well-defined doping
concentration. Considering also that p-type SnSe can be syn-
thesized by a very similar method,[11d] the way is clear toward
a unified, cost-effective processing route to large quantities
of both the constituent materials needed for a thermoelectric
device.
Experimental Section
Full experimental details are provided in the Supporting Information.
Materials Synthesis: 260 mmol citric acid and 10 mmol SnCl2·2H2O
were added into 50 mL deionized water (DIW) to yield a transparent
solution that was heated to boil. 50 mL of freshly prepared NaHSe(aq)
was promptly injected into the boiling solution. The solution was boiled
for 2 h and cooled to room temperature under Ar(g) on a Schlenk line.
The products were washed with DIW and ethanol and dried at 50 °C
for 12 h. Scaled-up syntheses were performed with 5.5-fold precursor
concentrations. The yield was 96(1)% of theoretical production. To tune
the particle size and Cl level, samples synthesized over 1 min, 5 min,
or 24 h durations were also prepared. For the surfactant-free synthesis,
4 mL hydrochloric acid was introduced into SnCl2 solution in place
of citric acid. Pellets were pressed in a graphite die under Ar (uniaxial
pressure of 60 MPa; 500 °C; 20 min).
Materials Characterization and Testing: PXD data were recorded
by a PANalytical X'pert Pro MPD diffractometer in Bragg-Brentano
geometry (Cu K
α
1 radiation,
λ
= 1.5406 Å). Rietveld refinement was
performed against PXD data using the GSAS and EXPGUI software
packages.[21] Imaging and elemental analysis were performed by SEM
(Carl Zeiss Sigma, at 5 and 20 kV, respectively) equipped with EDS
(Oxford Instruments X-Max 80). Further imaging, elemental analysis
and SAED were conducted by TEM (FEI Titan Themis 200 and JEOL
JEM-2011, operated at 200 kV). Optical bandgaps were measured by
DR-UV–Vis spectroscopy (Shimadzu, UV-2600). The Seebeck coefficient
and electrical conductivity of pellets were measured using a Linseis
LSR-3 instrument from 300 to 540 K. Thermal diffusivity (D) of pellets
was measured using a Linseis LFA 1000 instrument within the same
temperature range and thermal conductivity (
κ
) was calculated using
κ
=
DCp
ρ
, where Cp and
ρ
are specific heat capacity and density, respectively.
Hall measurements were performed on a nanometrics HL5500 Hall
system using a Van der Pauw configuration. The XPS experiments were
performed using a Kratos Axis Ultra-DLD photoelectron spectrometer
with an Al monochromatic X-ray source.
Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.
Acknowledgements
This work was financially supported by the EPSRC (EP/K022156/1).
The authors thank Peter Chung for assistance with SEM and Jialu
Chen for assistance with TEM elemental mapping. S.R.P. and J.-W.G.B.
acknowledge the EPSRC for support (EP/N01717X/1). H.F.G. and W.Z.
acknowledge the EPSRC for the Equipment Grant to purchase Titan
Themis 200 microscope (EP/L017008/1).
Received: October 21, 2016
Revised: January 5, 2017
Published online: February 20, 2017
[1] a) X. F. Duan, Y. Huang, Y. Cui, J. F. Wang, C. M. Lieber, Nature
2001, 409, 66; b) S. J. Oh, C. Uswachoke, T. S. Zhao, J. H. Choi,
B. T. Diroll, C. B. Murray, C. R. Kagan, ACS Nano 2015, 9, 7536.
[2] D. Zhitomirsky, M. Furukawa, J. Tang, P. Stadler, S. Hoogland,
O. Voznyy, H. Liu, E. H. Sargent, Adv. Mater. 2012, 24, 6181.
[3] a) J.-S. Rhyee, K. Ahn, K. H. Lee, H. S. Ji, J.-H. Shim, Adv. Mater.
2011, 23, 2191; b) G. Han, Z.-G. Chen, J. Drennan, J. Zou, Small
2014, 10, 2747.
[4] a) L. D. Zhao, S. H. Lo, J. Q. He, H. Li, K. Biswas, J. Androulakis,
C. I. Wu, T. P. Hogan, D. Y. Chung, V. P. Dravid, M. G. Kanatzidis,
J. Am. Chem. Soc. 2011, 133, 20476; b) Q. Zhang, Y. C. Lan,
S. L. Yang, F. Cao, M. L. Yao, C. Opeil, D. Broido, G. Chen,
Z. F. Ren, Nano Energy 2013, 2, 1121; c) H. Wang, Y. Z. Pei,
A. D. LaLonde, G. J. Snyder, Proc. Natl. Acad. Sci. USA 2012, 109,
9705; d) A. D. LaLonde, Y. Z. Pei, G. J. Snyder, Energy Environ. Sci.
2011, 4, 2090.
[5] H. Fang, Z. Luo, H. Yang, Y. Wu, Nano Lett. 2014, 14, 1153.
[6] J. Tang, K. W. Kemp, S. Hoogland, K. S. Jeong, H. Liu, L. Levina,
M. Furukawa, X. H. Wang, R. Debnath, D. K. Cha, K. W. Chou,
A. Fischer, A. Amassian, J. B. Asbury, E. H. Sargent, Nat. Mater.
2011, 10, 765.
[7] a) J. B. Zhang, J. B. Gao, E. M. Miller, J. M. Luther, M. C. Beard, ACS
Nano 2014, 8, 614; b) Z. Zhang, C. Liu, X. J. Zhao, J. Phys. Chem. C
2015, 119, 5626.
[8] a) M. Ibáñez, R. J. Korkosz, Z. S. Luo, P. Riba, D. Cadavid,
S. Ortega, A. Cabot, M. G. Kanatzidis, J. Am. Chem. Soc. 2015, 137,
CommuniCation
© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim (7 of 7) 1602328
wileyonlinelibrary.com
4046; b) D. James, X. Lu, A. C. Nguyen, D. Morelli, S. L. Brock,
J. Phys. Chem. C 2015, 119, 4635.
[9] a) G. J. Tan, L. D. Zhao, M. G. Kanatzidis, Chem. Rev. 2016, 116,
12123; b) W. G. Zeier, A. Zevalkink, Z. M. Gibbs, G. Hautier,
M. G. Kanatzidis, G. J. Snyder, Angew. Chem., Int. Ed. 2016, 55,
6826; c) G. J. Snyder, E. S. Toberer, Nat. Mater. 2008, 7, 105.
[10] a) L. D. Zhao, S. H. Lo, Y. S. Zhang, H. Sun, G. J. Tan, C. Uher,
C. Wolverton, V. P. Dravid, M. G. Kanatzidis, Nature 2014, 508,
373; b) L.-D. Zhao, G. Tan, S. Hao, J. He, Y. Pei, H. Chi, H. Wang,
S. Gong, H. Xu, V. P. Dravid, C. Uher, G. J. Snyder, C. Wolverton,
M. G. Kanatzidis, Science 2015, 351, 141; c) K. Peng, X. Lu, H. Zhan,
S. Hui, X. Tang, G. Wang, J. Dai, C. Uher, G. Wang, X. Zhou, Energy
Environ. Sci. 2016, 9, 454.
[11] a) E. K. Chere, Q. Zhang, K. Dahal, F. Cao, J. Mao, Z. Ren, J. Mater.
Chem. A 2016, 4, 1848; b) T.-R. Wei, G. Tan, X. Zhang, C.-F. Wu,
J.-F. Li, V. P. Dravid, G. J. Snyder, M. G. Kanatzidis, J. Am. Chem.
Soc. 2016, 138, 8875; c) C. L. Chen, H. Wang, Y. Y. Chen, T. Day,
G. J. Snyder, J. Mater. Chem. A 2014, 2, 11171; d) G. Han,
S. R. Popuri, H. F. Greer, J. W. G. Bos, W. Z. Zhou, A. R. Knox,
A. Montecucco, J. Siviter, E. A. Man, M. Macauley, D. J. Paul,
W. G. Li, M. C. Paul, M. Gao, T. Sweet, R. Freer, F. Azough, H. Baig,
N. Sellami, T. K. Mallick, D. H. Gregory, Angew. Chem., Int. Ed. 2016,
55, 6433; e) S. R. Popuri, M. Pollet, R. Decourt, F. D. Morrison,
N. S. Bennett, J. W. G. Bos, J. Mater. Chem. C 2016, 4, 1685;
f) Y.-X. Chen, Z.-H. Ge, M. Yin, D. Feng, X.-Q. Huang, W. Zhao,
J. He, Adv. Funct. Mater. 2016, 26, 6836.
[12] a) Q. Zhang, E. K. Chere, J. Y. Sun, F. Cao, K. Dahal, S. Chen,
G. Chen, Z. F. Ren, Adv. Energy Mater. 2015, 5, 1500360; b) X. Wang,
J. T. Xu, G. Q. Liu, Y. J. Fu, Z. Liu, X. J. Tan, H. Z. Shao, H. C. Jiang,
T. Y. Tan, J. Jiang, Appl. Phys. Lett. 2016, 108, 083902.
[13] PDF-2 Release 2008, Joint Committee on Powder Diffraction Stand-
ards (JCPDS)-International Centre for Diffraction Data (ICDD),
2008.
[14] D. D. Vaughn, S. I. In, R. E. Schaak, ACS Nano 2011, 5, 8852.
[15] B. Z. Sun, Z. J. Ma, C. He, K. C. Wu, Phys. Chem. Chem. Phys. 2015,
17, 29844.
[16] S. I. Kim, S. Hwang, S. Y. Kim, W. J. Lee, D. W. Jung, K. S. Moon,
H. J. Park, Y. J. Cho, Y. H. Cho, J. H. Kim, D. J. Yun, K. H. Lee,
I. T. Han, K. Lee, Y. Sohn, Sci. Rep. 2016, 6, 19733.
[17] D. Feng, Z.-H. Ge, D. Wu, Y.-X. Chen, T. Wu, J. Li, J. He, Phys. Chem.
Chem. Phys. 2016, 18, 31821.
[18] a) T. T. X. Vo, T. N. H. Le, Q. N. Pham, C. Byl, D. Dragoe,
M. G. Barthes-Labrousse, D. Berardan, N. Dragoe, Phys. Status
Solidi A 2015, 212, 2776; b) K. Rubenis, S. Populoh, P. Thiel,
S. Yoon, U. Müller, J. Locs, J. Alloys Compd. 2017, 692, 515.
[19] D. Martinez-Escobar, M. Ramachandran, A. Sanchez-Juarez,
J. S. N. Rios, Thin Solid Films 2013, 535, 390.
[20] S. Saha, A. Banik, K. Biswas, Chem. - Eur. J. 2016, 22, 15634.
[21] a) A. C. Larson, R. B. Von Dreele, General Structure Analysis System
(GSAS), Los Alamos National Laboratory Report LAUR 86-748, Los
Alamos National Laboratory, Los Alamos, NM 1994; b) B. H. Toby,
J. Appl. Crystallogr. 2001, 34, 210.
Adv. Energy Mater. 2017, 7, 1602328
www.advenergymat.dewww.advancedsciencenews.com
... PbS, [7] Bi 2 Te 3 , [40] Ag 2 Se [8] and SnSe. [8,15,41] Despite its importance and extended use on different systems, detailed studies in many of the reported systems is missing. ...
Article
Full-text available
Production of thermoelectric materials from solution‐processed particles involves the synthesis of particles, their purification and densification into pelletized material. Chemical changes that occur during each one of these steps render them performance determining. Particularly the purification steps, bypassed in conventional solid‐state synthesis, are the cause for large discrepancies among similar solution‐processed materials. In present work, the investigation focuses on a water‐based surfactant free solution synthesis of SnSe, a highly relevant thermoelectric material. We show and rationalize that the number of leaching steps, purification solvent, annealing, and annealing atmosphere have significant influence on the Sn : Se ratio and impurity content in the powder. Such compositional changes that are undetectable by conventional characterization techniques lead to distinct consolidated materials with different types and concentration of defects. Additionally, the profound effect on their transport properties is demonstrated. We emphasize that understanding the chemistry and identifying key chemical species and their role throughout the process is paramount for optimizing material performance. Furthermore, we aim to demonstrate the necessity of comprehensive reporting of these steps as a standard practice to ensure material reproducibility.
... Nanomaterials possess distinct characteristics that largely depend on their crystal phase, surface area, morphology and architecture [30,[67][68][69]. In contrast to bulk materials, nanomaterials exhibit size-and shape-dependent high-pressure phase transition behavior, including anomalous pressure responses and novel physical-chemical properties. ...
Article
Full-text available
Tin selenide (SnSe) holds great potential for abundant future applications, due to its exceptional properties and distinctive layered structure, which can be modified using a variety of techniques. One of the many tuning techniques is pressure manipulating using the diamond anvil cell (DAC), which is a very efficient in situ and reversible approach for modulating the structure and physical properties of SnSe. We briefly summarize the advantages and challenges of experimental study using DAC in this review, then introduce the recent progress and achievements of the pressure-induced structure and performance of SnSe, especially including the influence of pressure on its crystal structure and optical, electronic, and thermoelectric properties. The overall goal of the review is to better understand the mechanics underlying pressure-induced phase transitions and to offer suggestions for properly designing a structural pattern to achieve or enhanced novel properties.
Article
Conductive metal-organic frameworks (c-MOFs) are promising thermoelectric (TE) materials due to their low thermal conductivity and tunable electronic properties. Theoretical results show that two dimensional semiconuctors have natural advantages in...
Article
Production of thermoelectric materials from solution‐processed particles involves the synthesis of particles, their purification and densification into pelletized material. Chemical changes that occur during each one of these steps render them performance determining. Particularly the purification steps, bypassed in conventional solid‐state synthesis, are the cause for large discrepancies among similar solution‐processed materials. In present work, the investigation focuses on a water‐based surfactant free solution synthesis of SnSe, a highly relevant thermoelectric material. We show and rationalize that the number of leaching steps, purification solvent, annealing, and annealing atmosphere have significant influence on the Sn:Se ratio and impurity content in the powder. Such compositional changes that are undetectable by conventional characterization techniques lead to distinct consolidated materials with different types and concentration of defects. Additionally, the profound effect on their transport properties is demonstrated. We emphasize that understanding the chemistry and identifying key chemical species and their role throughout the process is paramount for optimizing material performance. Furthermore, we aim to demonstrate the necessity of comprehensive reporting of these steps as a standard practice to ensure material reproducibility.
Article
In the face of ever‐evolving energy and environmental challenges, tin selenide (SnSe) has garnered significant attention due to its outstanding thermoelectric performance. The article focuses on the research of easily accessible and highly practical polycrystalline SnSe thermoelectric materials, providing an overview of their crystal structure, band structure, and electrical transport performance. Compared with previous studies, this research classifies elements based on their own properties, mainly dividing them into alkali metals, transition metals, main group metals, halogens, and rare earth (Re) elements. The study systematically summarizes the experimental results of doping SnSe with these elements and analyzes the mechanisms by which different elements enhance the electrical transport performance and thermoelectric figure of merit of polycrystalline SnSe. The enhanced mechanism is mainly achieved by increasing the conductivity and Seebeck coefficient. Finally, a systematic analysis is conducted to identify the factors that improve electrical transport performance, and strategies for enhancing the electrical transport and thermoelectric properties of polycrystalline SnSe through precise doping techniques are discussed, with a prospect for their application in the future.
Article
Full-text available
An expansive effort has been made to optimize the thermoelectric performance of tin selenide (SnSe) due to its potential for waste heat recovery. p-Type co-doped SnSe with various doping percentages of copper (Cu) and silver (Ag) was synthesized hydrothermally to optimize the thermoelectric parameters. Temperature variation measurements of the thermoelectric parameters viz. Seebeck coefficient (S), electrical conductivity (σ), and thermal conductivity (κ) were performed for pristine, Cu-doped, and Cu/Ag codoped SnSe samples with a variation of the dopant percentages. Maximum S = 1225 μV K⁻¹ was obtained for pristine SnSe at a temperature of 373 K. The doped samples, comparatively, yielded lower S but higher σ coupled with lower κ. XPS results confirm the incorporation of Ag in the SnSe lattice with 0/+1 valence state, which is evidence for reduced thermal conductivity and enhanced electrical conductivity for Ag codoping in Cu-doped SnSe. DFT calculations have revealed the occurrence of point defects due to lattice distortion at the Ag and Cu dopant sites resulting in lattice anharmonicity-induced improved phonon scattering. Phonon and thermodynamic calculations also demonstrated a drastic decrease of 16.1% of specific heat capacity (Cp) for the Ag and Cu co-doped SnSe compared to pristine SnSe, along with a remarkable decrease in electron effective mass, group velocity at the BZ boundaries and phonon mode softening, thereby unraveling the origin of the improved performance. Maximum ZT = 0.54 was obtained for the SSCA-7-2 (SnSe + 7% Cu + 2% Ag) sample at 383 K.
Article
Full-text available
We present in this manuscript that enhanced thermoelectric performance can be achieved in polycrystalline SnSe prepared by hydrothermal reaction and spark plasma sintering (SPS). X-ray diffraction (XRD) patterns revealed strong orientation along the [l 0 0] direction in bulk samples, which was further confirmed by microstructural observation through transmission electron microscopy (TEM) and field emission scanning electron microscopy (FESEM). It was noticed that the texturing degree of bulk samples could be controlled by sintering temperature during the SPS process. The best electrical transport properties were found in the sample which sintered at 450 °C in the direction vertical to the pressing direction, where the highest texturing degree and mass density were achieved. Coupled with the relatively low thermal conductivity, an average ZT of ∼ 0.38, the highest ever reported in pristine polycrystalline SnSe was obtained. This work set up a forceful example that a texture-control approach can be utilized to enhance the thermoelectric performance effectively.
Article
Full-text available
A surfactant-free solution methodology, simply using water as a solvent, has been developed for the straightforward synthesis of single-phase orthorhombic SnSe nanoplates in gram quantities. Individual nanoplates are composed of {100} surfaces with {011} edge facets. Hot-pressed nanostructured compacts (Eg ≈0.85 eV) exhibit excellent electrical conductivity and thermoelectric power factors (S(2) σ) at 550 K. S(2) σ values are 8-fold higher than equivalent materials prepared using citric acid as a structure-directing agent, and electrical properties are comparable to the best-performing, extrinsically doped p-type polycrystalline tin selenides. The method offers an energy-efficient, rapid route to p-type SnSe nanostructures.
Article
In the present study thermoelectric properties of spark plasma sintered and subsequently air-annealed Sn1-xSbxO2 (x = 0, 0.01, 0.03, 0.05) ceramic samples were characterized. It was observed that addition of Sb greatly affects density, microstructure and thermoelectric properties of SnO2. Grain size decreases and the microstructure becomes more homogeneous with increasing Sb content in the Sn1-xSbxO2 samples from x = 0.01 to 0.05. The thermal conductivity and Seebeck coefficient of the Sb doped samples decreases with an increase in Sb concentration, while electrical conductivity increases up to a maximum at Sb doping level of x = 0.03 and then decreases at x = 0.05. The highest thermoelectric figure of merit of ZT ≈ 0.06 at 1073K was exhibited by the sample of composition Sn0.99Sb0.01O2.
Article
Layered p-block metal chalcogenides are renowned for thermoelectric energy conversion due to their low thermal conductivity caused by bonding asymmetry and anharmonicity. Recently, single crystalline layered SnSe has created sensation in thermoelectrics due to its ultralow thermal conductivity and high thermoelectric figure of merit. Tin diselenide (SnSe2) is an additional layered compound belonging to Sn-Se phase diagram, which possesses CdI2 type structure. However, synthesis of pure phase bulk SnSe2 by conventional solid state route is still remains challenging. Herein, we present a simple solution-based low-temperature synthesis of ultrathin (3-5 nm) few layers (4-6 layers) nanosheets of Cl doped SnSe2, which possess n-typecarrier concentration of 2 × 1018 cm3 with carrier mobility of ~30 cm2V-1s-1 at room temperature. SnSe2 exhibits band gap of ~1.6 eV and semiconducting electronic transport in the 300-630 K range. An ultralow thermal conductivity of ~0.67 Wm-1K-1 has been achieved at room temperature in hot-pressed dense pellet of Cl doped SnSe2 nanosheets due to anisotropic layered structure, which gives rise to effective phonon scattering.
Article
There has been a renaissance of interest in exploring highly efficient thermoelectric materials as a possible route to address the worldwide energy generation, utilization, and management. This review describes the recent advances in designing high-performance bulk thermoelectric materials. We begin with the fundamental stratagem of achieving the greatest thermoelectric figure of merit ZT of a given material by carrier concentration engineering, including Fermi level regulation and optimum carrier density stabilization. We proceed to discuss ways of maximizing ZT at a constant doping level, such as increase of band degeneracy (crystal structure symmetry, band convergence), enhancement of band effective mass (resonant levels, band flattening), improvement of carrier mobility (modulation doping, texturing), and decrease of lattice thermal conductivity (synergistic alloying, second-phase nanostructuring, mesostructuring, and all-length-scale hierarchical architectures). We then highlight the decoupling of the electron and phonon transport through coherent interface, matrix/precipitate electronic bands alignment, and compositionally alloyed nanostructures. Finally, recent discoveries of new compounds with intrinsically low thermal conductivity are summarized, where SnSe, BiCuSeO, MgAgSb, complex copper and bismuth chalcogenides, pnicogen-group chalcogenides with lone-pair electrons, and tetrahedrites are given particular emphasis. Future possible strategies for further enhancing ZT are considered at the end of this review.
Article
P-type polycrystalline SnSe and K0.01Sn0.99Se are prepared by combining mechanical alloying (MA) and spark plasma sintering (SPS). The highest ZT of ≈0.65 is obtained at 773 K for undoped SnSe by optimizing the MA time. To enhance the electrical transport properties of SnSe, K is selected as an effective dopant. It is found that the maximal power factor can be enhanced significantly from ≈280 μW m−1 K−2 for undoped SnSe to ≈350 μW m−1 K−2 for K-doped SnSe. It is also observed that the thermal conductivity of polycrystalline SnSe can be enhanced if the SnSe powders are slightly oxidized. Surprisingly, after K doping, the absence of Sn oxides at grain boundaries and the presence of coherent nanoprecipitates in the SnSe matrix contribute to an impressively low lattice thermal conductivity of ≈0.20 W m−1 K−1 at 773 K along the sample section perpendicular to pressing direction of SPS. This extremely low lattice thermal conductivity coupled with the enhanced power factor results in a record high ZT of ≈1.1 at 773 K along this direction in polycrystalline SnSe.
Article
Recent findings about ultrahigh thermoelectric performance in SnSe single crystals have stimulated the related researches on this simple binary compound, which are most devoted to its polycrystalline counterparts with a focus on electrical property enhancement by effective doping. This work systematically investigated the thermoelectric proper-ties of polycrystalline SnSe doped with three alkali metals (Li, Na and K). It is found that Na has the best doping effi-ciency, leading to an increase in hole concentration from 3.2×1017 to 4.4×1019 cm-3 at room temperature, accompanied with a drop in Seebeck coefficient from 480 to 142 μV/K. An equivalent single parabolic band model was found ade-quate to capture the variation tendency of Seebeck coefficient with doping levels within a wide range. A mixed scattering of carriers by acoustic phonons and grain boundaries is suitable for numerically understanding the temperature-dependence of carrier mobility. A maximum ZT of ~0.8 was achieved in 1% Na- or K-doped SnSe at 800 K. Possible strategies to improve the mobility and ZT of polycrystals were also proposed.
Article
The coupled transport properties required to create an efficient thermoelectric material necessitates a thorough understanding of the relationship between the chemistry and physics in a solid. We approach thermoelectric material design using the chemical intuition provided by molecular orbital diagrams, tight binding theory, and a classic understanding of bond strength. Concepts such as electronegativity, band width, orbital overlap, bond energy, and bond length are used to explain trends in electronic properties such as the magnitude and temperature dependence of band gap, carrier effective mass, and band degeneracy and convergence. The lattice thermal conductivity is discussed in relation to the crystal structure and bond strength, with emphasis on the importance of bond length. We provide an overview of how symmetry and bonding strength affect electron and phonon transport in solids, and how altering these properties may be used in strategies to improve thermoelectric performance.
Article
N-type SnSe compound has been synthesized through melting with spark plasma sintering. By doping BiCl3, the carrier concentration of SnSe is significantly increased, leading to a large enhancement of electrical conductivity. Meanwhile, the SnSe0.95-BiCl3 samples also exhibit higher Seebeck coefficient and lower lattice thermal conductivity, compared with polycrystalline SnSe. Consequently, a high power factor of ∼5 μW cm−1 K−2 and a ZT of 0.7 have been achieved at 793 K. The synergistic roles of BiCl3doping in SnSe provide many opportunities in the optimization of n-type SnSe materials.
Article
Single crystals of SnSe have been reported to have very high thermoelectric efficiencies with a maximum figure merit zT = 2.5. This outstanding performance is due to ultralow thermal conductivities. We report on the synthesis of highly textured polycrystalline SnSe ingots with large single-crystal magnitude power factors, S2/ρ = 0.2-0.4 mW m-1 K-2 between 300-600 K, increasing to 0.9 mW m-1 K-2 at 800 K, and bulk thermal conductivity values κ300K = 1.5 W m-1 K-1. However, small SnSe ingots, which were measured in their entirety, were found to have a substantially reduced κ300K = 0.6 W m-1 K-1. Microscopy and diffraction revealed two distinct types of texturing within the hot-pressed ingots. In the interior, large coherent domains of SnSe platelets with a ~45° orientation with respect to the pressing direction are found, while the platelets are preferentially oriented at 90° to the pressing direction at the top and bottom of the ingots. Fitting the κ(T) data suggests an increase in defect scattering for the smaller ingots, which is in keeping with the presence of regions of structural disorder due to the change in texturing. Combining the measured S2/ρ with the bulk ingot κ values yields zT = 1.1 at 873 K.