ArticlePDF Available

Facile Surfactant-Free Synthesis of p-Type SnSe Nanoplates with Exceptional Thermoelectric Power Factors

Wiley
Angewandte Chemie International Edition
Authors:

Abstract and Figures

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.
Content may be subject to copyright.
German Edition:DOI:10.1002/ange.201601420
Thermoelectrics Very Important Paper International Edition:DOI:10.1002/anie.201601420
Facile Surfactant-Free Synthesis of p-Type SnSe Nanoplates with
ExceptionalThermoelectric PowerFactors
Guang Han, Srinivas R. Popuri, Heather F. Greer,Jan-Willem G. Bos,Wuzong Zhou,
Andrew R. Knox, Andrea Montecucco,Jonathan Siviter,Elena A. Man, Martin Macauley,
Douglas J. Paul, Wen-guang Li, Manosh C. Paul, Min Gao,Tracy Sweet, Robert Freer,
Feridoon Azough, Hasan Baig,Nazmi Sellami, Tapas K. Mallick, and Duncan H. Gregory*
Abstract: Asurfactant-free solution methodology,simply
using water as asolvent, has been developed for the straight-
forwardsynthesis of single-phase orthorhombic SnSe nano-
plates in gram quantities.Individual nanoplates are composed
of {100} surfaces with {011} edge facets.Hot-pressed nano-
structured compacts (Eg0.85 eV) exhibit excellent electrical
conductivity and thermoelectric power factors (S2s)at550 K.
S2svalues are 8-fold higher than equivalent materials prepared
using citric acid as astructure-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.
Growing global energy demands,together with the negative
impacts resulting from combustion of fossil fuels,have
diverted attention to technologies for sustainable energy
generation and conversion.[1] Thermoelectrics realize direct
inter-conversion between thermal and electrical energy and
provide opportunities to harvest useful electricity from waste
heat (and conversely to perform refrigeration). Thethermo-
electric conversion efficiencyofamaterial is determined by
its dimensionless figure of merit, ZT=S2sT/k,where S,s,T,
and krepresent the Seebeck coefficient, electrical conductiv-
ity,absolute temperature,and thermal conductivity,respec-
tively.[2] Extensive efforts have been devoted to the improve-
ment of the thermoelectric performance of state-of-the-art
materials,[3] and to the discovery of new thermoelectrics[4]
with ZT values >2. Single-crystalline SnSe combines ahigh
ZTwith arelatively low toxicity and high Earth-abundance of
the component elements.[4] SnSe crystals possess very low
thermal conductivity owing to lattice anharmocity,yielding
record high ZTvalues of 2.6 and 2.3 at 923 Kalong the band c
crystallographic directions,respectively.[4] Polycrystalline
SnSe materials have been fabricated to improve mechanical
properties,[5] but ZT has been limited to 1, owing to both
increased electrical resistivity and thermal conductivity.[5]
Unfortunately,the synthesis of SnSe is protracted and
energy-intensive,involving heating, melting, and annealing
at high temperatures ( 800–1223 K).[4–5] Before the potential
of SnSe can be realized, afast, cost-effective,and large-scale
synthesis route to the pure selenide that does not sacrifice
performance is essential.
Nanostructuring very effectively enhances ZT. Thehigh
density of interfaces improves phonon scattering, decreasing
the lattice thermal conductivity.[2, 3] Bottom-up solution syn-
thesis methods facilitate control of size,morphology,crystal
structure,and defects.[6] However,the organic surfactants that
can control morphology through surface modification are
commonly electrically insulating, which can drastically reduce
the electrical conductivity of the materials.[7] Ligand replace-
ment methods switch smaller species for long chain surfactant
molecules,[7] but sometimes involve using high toxicity
chemicals,[8] and introduce impurities,[7b] which again can
adversely influence the transport behavior of the materials.[7b]
Organic contamination can be prevented only if asuitable
surfactant-free synthesis strategy can be found,[9] and to date
solution syntheses of SnSe nanostructures have required
organic surfactants and/or solvents,for example,oleyamine,
[*] Dr.G.Han, Prof. D. H. Gregory
WestCHEM, SchoolofChemistry, 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 &Recovery
School of Engineering &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)
Prof. A. R. Knox, Dr.A.Montecucco, Dr.J.Siviter,Dr. E. A. Man,
Dr.M.Macauley,Prof. D. J. Paul, Prof. W.-g. Li, Dr.M.C.Paul
School of Engineering, University of Glasgow
Glasgow,G12 8QQ (UK)
Dr.M.Gao, Dr.T.Sweet
School of Engineering, Cardiff University
Cardiff, CF24 3AA (UK)
Prof. R. Freer,Dr. F. Azough
Materials Science Centre, School of Materials
UniversityofManchester
Manchester,M13 9PL (UK)
Dr.H.Baig, Dr.N.Sellami, Prof. T. K. Mallick
Environmentand Sustainability Institute, University of Exeter
Penryn Campus, Penryn TR10 9FE (UK)
Supportinginformation and the ORCID identification number(s) for
the author(s) of this article can be found under
http://dx.doi.org/10.1002/anie.201601420.
A
ngewandte
Chemie
Communications
6433Angew.Chem. Int.Ed. 2016,55,6433 –6437 Ó2016Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim
trioctylphosphine selenide,and bis[bis(trimethylsilyl)
amino]tin(II), while only yielding milligram quantities of
materials.[10] In this study,wedemonstrate asurfactant-free
aqueous solution approach towards the preparation of >10 g
of SnSe nanoplates,byboiling amixture of NaHSe and
Na2SnO2solutions for 2h.The phase-pure nanoplates can be
hot pressed into dense pellets with outstanding thermoelectric
power factors (Scheme 1).
Injecting NaHSe into aNa
2SnO2solution leads to the
instant precipitation of SnSe nanoparticles [Supporting Infor-
mation, Figure S1;Eq. (1)]:
NaHSe þNa2SnO2þH2O!SnSe þ3NaOH ð1Þ
Boiling the suspension for 2hleads to the formation of
crystalline,phase-pure nanoplates of orthorhombic SnSe
(ICDD card No.48-1224).[11] Rietveld refinement against
powder X-ray diffraction (PXD) data (Figure 1a;
Tables S1,S2) confirmed the orthorhombic structure (space
group Pnma,a=11.5156(5), b=4.1571(2), c=4.4302(3) è).
Scanning electron microscopy (SEM;Figures 1b,S2a)
revealed that the product is comprised of rectangular nano-
plates,each with lateral dimensions of 80–200 nm and athick-
ness of 10–60 nm. Energy dispersive X-ray spectroscopy
(EDS;Figure S2b) consistently generated Sn:Se ratios of
49(1):51(1) atom%, while Fourier transform infrared (FTIR)
spectra (Figure S3) provided no evidence for organic func-
tional groups (as might be expected from equivalent surfac-
tant-assisted syntheses).
Transmission electron microscopy (TEM;Figure 1c)
showed that the SnSe nanoplates were almost uniformly
rectangular,and selected area electron diffraction (SAED)
patterns obtained with the incident electron beam normal to
the face of the nanoplate could be indexed along the [100]
SnSe zone axis.Aset of lattice spacings of 3.0 èintersect-
ing with an angle of 93(1)88could be measured from high
resolution TEM (HRTEM;Figure 1d)corresponding to the
{011} plane spacings.Combined with SAED data, the nano-
plate face can thus be identified as the bc plane of SnSe and
the side facets are defined by {011} planes (Figures 1c,S4).
Theobserved splitting in diffraction spots suggested twin
defects induced by orthorhombic distortion.[12] Images and
SAED patterns along the [001] zone axis (beam direction
parallel to the nanoplate face;Figure 1e)verified that :i)the
plates are approximately an order of magnitude thinner in the
third dimension, and ii)the bc plane forms the nanoplate
faces.Further,diffraction spots are elongated along [100],
indicating planar defects along the aaxis.[13] Lattice spacings
of 5.7 è(d(200))and 4.2 è(d(010))were observed in the
corresponding HRTEM image (Figure 1f).
Intermediate products synthesized after only 1min of
heating were investigated to understand the morphological
evolution. Theproduct is single-phase orthorhombic SnSe
(Figure S5a) composed principally of irregular, near-rectan-
gular nanoplates,many of which are truncated (Figure S5b).
TEM revealed internal angles of 133(1)88and 94(1)88at the
truncated and regular corners,respectively (Figure 2a). When
Scheme 1. The solution synthesis and hot pressing of SnSe nano-
plates:a)injection of NaHSe(aq) into Na2SnO2(aq) to trigger the
reaction;b)formation of SnSe nanoplates;c)orientation of nanoplates
induced by hot pressing;d)structure model of fabricated bulk pellets.
The insets in (b) show the nanoplate solution and the yield ( 94 %)
from a2hsynthesis.
Figure 1. Characterization of SnSe nanoplates synthesized after 2h:
a) profile plot from Rietveld refinement; b) SEM image ;c)TEM image
of aSnSe nanoplate and its corresponding SAED pattern along the
[100] zone axis;d)HRTEM image of part of the plate shown in (c)
with d-spacings indicated;e)profile HRTEM image of aSnSe nano-
plate and its corresponding SAED pattern along the [001] zone axis;
and f) HRTEM image of part of the plate shown in (e) with d-spacings
indicated.
A
ngewandte
Chemie
Communications
6434 www.angewandte.org Ó2016 Wiley-VCH Verlag GmbH &Co. KGaA,Weinheim Angew.Chem.Int. Ed. 2016,55,6433 –6437
correlating the SAED pattern (along the [100] zone axis;
Figure 2b), the TEM image and the crystal plane intersection
angle along the bc plane (Figure S4) the facets of the SnSe
truncated nanoplate can be depicted as shown in Figure 2a.
Hence,the SnSe truncated nanoplate is enclosed by {100} and
{011}, together with {001} facets.Given that no surfactant is
used, the nanoplate shape is determined primarily by the
intrinsic features of the anisotropic selenide crystal structure.
Atomic planes with high surface energies usually exhibit fast
growth rates,and in SnSe the {001} and {010} planes possess
much higher surface energies than the {011} planes.[14] The
former planes would thus experience faster initial growth
than the {011} planes.Tomaintain the minimum surface
energy as growth progresses,the {001} and {010} planes
diminish, while the {011} planes feature increasingly in the
side facets (Figures 2a,c) until they dominate completely
(Figures 1c,2d). TheNaOH concentration is also important
in regulating growth, and by decreasing the molar ratio from
15:1 to 15:2 the mean length/width of the SnSe nanoplates is
reduced from 150 nm to 80 nm (Figures S6, S7). Decreas-
ing the hydroxide concentration further has more profound
effects on the reaction chemistry (see the Supporting
Information).
Theability to prepare >10 gsurfactant-free SnSe nano-
materials allowed the fabrication of high-density pellets
through hot pressing without the necessity of high temper-
ature annealing. Pellets of 95%theoretical density,retain-
ing the orthorhombic SnSe structure were obtained (denoted
1;Figure 3a;Tables S3, S4). Strong orientation of the plates in
the bc plane is evidenced by the increased intensity of the
(h00) PXD reflections,and the decrease in peak half-widths
indicates alarger crystallite size after hot pressing.The
indirect (direct) optical band gap from diffuse reflectance
(DR) UV/Vis spectra[10c] narrows slightly from 0.89
(1.1) eV to 0.85 ( 1.0) eV (Figure S10) when the nano-
plates are consolidated into dense pellets,which could be
related to sintering effects.The values are very similar to the
indirect band gaps reported for both single crystalline and
polycrystalline SnSe.[4,5c, d] 1is composed of densely packed
particles,typically 200 nm across with flat surfaces (Fig-
ures 3b,S11a). TheSn:Se ratio remains at 49(1):51(1) atom%
(Figure S11b). An SAED pattern (Figure 3c), with the beam
normal to the face of ananoplate taken from 1was indexed
along the SnSe [100] zone axis.The single-crystal structure
was confirmed by the HRTEM image (Figure 3d). TEM also
showed that the nanoplate from 1consisted of compacted
smaller platelets (Figure S11c). Thermogravimetric analysis
(TGA) of 1under both argon and air revealed negligible
weight changes below 50088C, but suggested that thermal
decomposition and oxidation, respectively,begin above this
temperature (Figure S12).
Forcomparison, asecond sample of SnSe nanoparticles
(40–60 nm) were synthesized by acitric-acid-assisted
solution synthesis,which were also consolidated into dense
pellets ( 92%ofthe theoretical density) by hot pressing
(denoted 2;Figures S13, S14). Compared to 1,2possesses the
same orthorhombic structure,asimilar optical band gap and
forms comparable nanostructures ( 200 nm across oriented
in the bc plane). Importantly,however, Cl is detected in 2
(Sn:Se :Cl ratios of 51(1):48(1):1(1) atom%) that likely
originates from the replacement of ligated citric acid by Cl
during processing.[7b] Thesimilar densities and constituent
particle sizes of 1and 2allowed for agood comparison of their
relative electrical performance.The electrical conductivity of
1(Figure 4a)increases four-fold from 840 Sm¢1at 300 Kto
3500 Sm¢1at 550 K. Themagnitude of the values for 1can
be attributed to the high crystallinity,small band gap,
surfactant-free particle surface,microstructural orientation,
and the high level of sintering and densification achieved. By
contrast, 2exhibited electrical conductivity increasing from
55 Sm¢1at 300 Ktoonly 250 Sm¢1at 550 K; more than
an order of magnitude lower than 1.
Figure 2. Characterization of SnSe nanostructuressynthesized after
1min:a,b)TEM image and corresponding SAED pattern along the
[100] zone axis of aSnSe truncatednanoplate; and c, d) structure
models of individual SnSe nanoplates with and without truncation,
respectively,established on the basis of the detailed TEM character-
ization.
Figure 3. Characterization of SnSe pellet 1:a)profile plot for 1from
Rietveld refinement against PXD data;b)SEM image of the surface of
1;c)TEM image of aSnSe peeled nanoplate and its corresponding
SAED pattern along the [100] zone axis from the circled area;
d) HRTEM image of part of the plate shown in (c) with d-spacings
indicated.
A
ngewandte
Chemie
Communications
6435Angew.Chem. Int.Ed. 2016,55,6433 –6437 Ó2016 Wiley-VCH Verlag GmbH &Co. KGaA, We inheim www.angewandte.org
Thecontrast in the variation in the Seebeck coefficient
with temperature for 1and 2is striking (Figure 4b). Sfor
1increases almost linearly with temperature (250 mVK¢1at
room temperature to 340 mVK¢1at 550 K). By comparison, 2
shows n-type behavior at room temperature (S
¢150 mVK¢1), with the value of Sbecoming positive (p-type
behaviour) at 490 Kand rising to 80 mVK¢1at 550 K. It is
possible that the n-type conducting behavior correlates to the
presence of Cl and/or aslight excess of Sn, as noted above.We
are currently investigating this behavior further in systematic
doping experiments.Ann/p or p/n inversion with increasing
temperature has also been observed in pellets consolidated
from PbTe, Ag2Te,and PbTe0.1Se0.4S0.5 synthesized through
surfactant-assisted solution methods,[7,15] and should be
related to the thermal activation of higher concentrations of
positive or negative charge carriers,respectively.[7b] It is also
notable that both sand Sincrease with temperature for 1.
This phenomenon has been observed in both un-doped and
iodine-doped polycrystalline SnSe.[5c,d,16] Although the origins
of the behavior for 1require further investigation, the
combination of superior svalues coupled with high values
of Sleads to exceptional power factors ( 0.05 mW m¢1K¢2at
300 Kto0.40 mWm¢1K¢2at 550 K; Figure 4c). In contrast,
the power factors for 2are much lower, (0.001 mW m¢1K¢2at
300 Kand reaching only 0.05 mW m¢1K¢2at 550 K). Thehuge
differences in performance between 1and 2further empha-
size the importance of the surfactant-free synthesis route,not
just in the context of asimpler, more sustainable synthesis
method, but also in delivering significantly improved elec-
trical properties consistently (Figure S15). Notably,the power
factors for 1far exceed those for un-doped polycrystalline
SnSe across asimilar temperature range (0.028–
0.04 mWm¢1K¢2),[5c–e] and are comparable to those for hole-
doped materials with high carrier concentrations.[5d,17] Recent
Na- and Ag-doping studies have elegantly demonstrated how
the electrical performance and ZT values of SnSe single
crystals can be dramatically improved.[18] Given that the
samples in our studies were non-optimized, strategies involv-
ing systematic hole doping,inconjunction with surfactant-
free nanostructuring approaches,should yield even higher
performing p-type SnSe materials and pave the way for one-
pot synthesis of p- and n-type SnSe nanomaterials.
In summary,asimple,quick, surfactant-free,and energy-
efficient solution synthesis yielded SnSe nanoplates in gram
quantities.The ensuing nanostructured pellets exhibited
exceptional electrical conductivity coupled with high Seebeck
coefficients,leading to power factors surpassing those of
polycrystalline and surfactant-coated counterparts.The tech-
nique should be readily adaptable to include dopants and
amenable to the discovery of further materials,both p- and n-
type,with enhanced thermoelectric properties.
Experimental Section
Full experimentaldetails are provided in the SupportingInformation.
Materials Synthesis.100 mmol NaOH and 10 mmol SnCl2·2H2O
were added into 50 mL deionizedwater to yield atransparent
Na2SnO2solution. 50 mL of NaHSe(aq) preparedfrom Se and NaBH4
was injected into the boiling solution, leading to the immediate
formationofablack precipitate.The mixture was boiled for 2h,and
cooled to room temperature under Ar(g) on aSchlenk line.The
products were washed with deionized water and ethanol and dried at
5088Cfor 12 h. Scaled-up syntheses were performed with six-fold
precursor concentrations (94(1)% yield). Forthe surfactant-assisted
synthesis,50gcitric acid was introduced into SnCl2solution with no
addition of NaOH and the reaction duration was increased to 24 h.
Materials Characterization and Testing.PXD was performed
using aPANalyticalXpert Pro MPD diffractometerinBragg–
Brentano geometry (Cu Ka1radiation, l=1.5406 è). Rietveld
refinementwas performedusing the GSAS and EXPGUI software
packages,[19] with the previously published SnSe structureasarefer-
ence.[20] Imaging and elemental analysis were performedbySEM
(Carl Zeiss Sigma, at 5and 20 kV respectively)equippedwith EDS
(Oxford Instruments X-Max 80). Further imaging and SAED was
conducted by TEM (JEOL 2011, operated at 200 kV). TheSeebeck
coefficient and electrical conductivityof1and 2were measured using
aLinseis LSR-3 instrument from 300–550 K. Pellets were pressed in
agraphite die under Ar (uniaxial pressure of 60 MPa;50088C;
20 min).
Acknowledgements
This work was financially supported by the EPSRC (EP/
K022156/1). Theauthors thank Peter Chung for assistance
with SEM. SRP and JWGB acknowledge the Leverhulme
Trust RPG 2012-576.
Keywords: nanomaterials ·structures ·synthesis ·
thermoelectrics ·tin selenide
Howtocite: Angew.Chem. Int. Ed. 2016,55,6433 –6437
Angew.Chem. 2016,128,6543–6547
[1] a) J. R. Szczech,J.M.Higgins,S.Jin, J. Mater.Chem. 2011,21,
4037 –4055 ;b)G.Han, Z.-G.Chen, J. Drennan, J. Zou, Small
2014,10,2747 –2765.
[2] a) M. S. Dresselhaus,G.Chen, M. Y. Tang,R.G.Yang,H.Lee,
D. Z. Wang,Z.F.Ren, J. P. Fleurial, P. Gogna, Adv.Mater. 2007,
19,1043 –1053;b)L.D.Zhao,V.P.Dravid, M. G. Kanatzidis,
Energy Environ. Sci. 2014,7,251 –268.
[3] K. Biswas,J.Q.He, I. D. Blum, C. I. Wu,T.P.Hogan, D. N.
Seidman,V.P.Dravid, M. G. Kanatzidis, Nature 2012,489,414
418.
[4] 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 –377.
[5] a) S. Sassi, C. Candolfi, J. B. Vaney,V.Ohorodniichuk, P.
Masschelein, A. Dauscher,B.Lenoir, Appl. Phys.Lett. 2014,
104,212105;b)C.L.Chen, H. Wang,Y.Y.Chen, T. Day,G.J.
Figure 4. Electrical properties of SnSe pellets 1and 2measured
perpendiculartothe hot pressing direction:a)the electrical conductiv-
ity (s), b) the Seebeck coefficient(S), and c) the power factor (S2s)as
afunction of temperature.
A
ngewandte
Chemie
Communications
6436 www.angewandte.org Ó2016 Wiley-VCH Verlag GmbH &Co. KGaA,Weinheim Angew.Chem.Int. Ed. 2016,55,6433 –6437
Snyder, J. Mater.Chem. A 2014,2,11171 –11176;c)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;d)T.R.Wei, C. F.
Wu,X.Z.Zhang,Q.Tan, L. Sun, Y. Pan, J. F. Li, Phys.Chem.
Chem. Phys. 2015,17,30102 –30109;e)Y.L.Li, X. Shi, D. D.
Ren, J. K. Chen, L. D. Chen, Energies 2015,8,6275 –6285 ;
f) S. R. Popuri, M. Pollet, R. Decourt, F. D. Morrison, N. S.
Bennett, J. W. G. Bos, J. Mater.Chem. C 2016,4,1685 –1691.
[6] R. J. Mehta, Y. L. Zhang,C.Karthik, B. Singh, R. W. Siegel, T.
Borca-Tasciuc,G.Ramanath, Nat. Mater. 2012,11,233 –240.
[7] a) D. Cadavid, M. IbÇez, A. Shavel, O. J. Dur, M. A.
Lýpez de laTorre,A.Cabot, J. Mater.Chem. A 2013,1,4864
4870;b)M.IbÇez, R. J. Korkosz, Z. S. Luo,P.Riba, D. Cadavid,
S. Ortega, A. Cabot, M. G. Kanatzidis, J. Am. Chem. Soc. 2015,
137,4046 –4049.
[8] F. J. Fan, Y. X. Wang,X.J.Liu, L. Wu,S.H.Yu, Adv.Mater.
2012,24,6158 –6163.
[9] a) Y. Min, J. W. Roh, H. Yang,M.Park, S. I. Kim, S. Hwang,
S. M. Lee,K.H.Lee,U.Jeong, Adv.Mater. 2013,25,1425 –1429;
b) Y. C. Zhang,H.Wang,S.Kraemer,Y.F.Shi, F. Zhang,M.
Snedaker, K. L. Ding,M.Moskovits,G.J.Snyder,G.D.Stucky,
ACSNano 2011,5,3158 –3165;c)C.Han, Z. Li, G. Q. Lu, S. X.
Dou, Nano Energy 2015,15,193 –204.
[10] a) W. J. Baumgardner,J.J.Choi, Y. F. Lim, T. Hanrath, J. Am.
Chem. Soc. 2010,132,9519 –9521;b)M.A.Franzman,C.W.
Schlenker, M. E. Thompson, R. L. Brutchey, J. Am. Chem. Soc.
2010,132,4060 –4061;c)D.D.Vaughn, S. I. In, R. E. Schaak,
ACSNano 2011,5,8852 –8860;d)L.Li, Z. Chen, Y. Hu, X. W.
Wang,T.Zhang,W.Chen, Q. B. Wang, J. Am. Chem. Soc. 2013,
135,1213 –1216.
[11] PDF-2 Release 2008, Joint Committee on Powder Diffraction
Standards(JCPDS)-International Centre for DiffractionData
(ICDD), 2008.
[12] Y. B. Wang,F.Guyot, R. C. Liebermann, J. Geophys.Res. 1992,
97,12327 –12347.
[13] J. P. Ye,S.Soeda, Y. Nakamura, O. Nittono, Jpn. J. Appl. Phys.
1998,37,4264 –4271.
[14] X. H. Ma, K. H. Cho,Y.M.Sung, CrystEngComm 2014,16,
5080 –5086.
[15] D. Cadavid, M. IbaÇez, S. Gorsse,A.M.Lýpez, A. Cirera, J. R.
Morante,A.Cabot, J. Nanopart. Res. 2012,14,1328.
[16] F. Serrano-Snchez, M. Gharsallah,N.M.Nemes,F.J.Mom-
pean, J. L. Martnez,J.A.Alonso, Appl. Phys.Lett. 2015,106,
083902.
[17] a) E. K. Chere,Q.Zhang,K.Dahal, F. Cao,J.Mao,Z.Ren, J.
Mater.Chem. A 2016,4,1848 –1854;b)H.-Q.Leng,M.Zhou, J.
Zhao,Y.-M. Han, L.-F.Li, RSC Adv. 2016,6,9112 –9116.
[18] a) K. Peng,X.Lu, H. Zhan, S. Hui, X. Ta ng,G.Wang,J.Dai, C.
Uher,G.Wang,X.Zhou, Energy Environ. Sci. 2016,9,454 –460;
b) L.-D.Zhao,G.Tan, S. Hao,J.He, Y. Pei, H. Chi, H. Wa ng,S.
Gong,H.Xu, V. P. Dravid, C. Uher,G.J.Snyder,C.Wolverton,
M. G. Kanatzidis, Science 2015,351,141 –144.
[19] a) A. C. Larson, R. B. Vo nDreele,General Structure Analysis
System (GSAS);Los Alamos National Laboratory Report
LAUR 86 –748;Los Alamos National Laboratory, 1994;
b) B. H. Toby, J. Appl. Crystallogr. 2001,34,210 –213.
[20] A. S. Avilov,R.M.Imamov, S. N. Navasardyan, Kristallografiya
1979,24,874 –875.
Received:February 9, 2016
Revised: March 23, 2016
Published online: April 20, 2016
A
ngewandte
Chemie
Communications
6437Angew.Chem. Int.Ed. 2016,55,6433 –6437 Ó2016 Wiley-VCH Verlag GmbH &Co. KGaA, We inheim www.angewandte.org
... Solution-based syntheses surged as a viable alternative to prepare TE materials from particles made in solution. [10,11,19,[28][29][30][31] Particle syntheses can be performed with or without surfactants (also called ligands), at temperatures ranging from room temperature to around 350°C, with reaction times spanning from a few minutes [32,33] to even hours. [8,15,34] Solvents employed range from water [7,8,35,36] to organic solvents [37] and molten salts, [38] in some cases under pressure, e.g. ...
... The water-based synthesis is also very versatile, allowing for modifications such as doping, to further tailor the TE performance. [31,39,40] Additionally, this synthetic approach has demonstrated to yield various TE materials with enhanced performance, e.g. PbS, [7] Bi 2 Te 3 , [40] Ag 2 Se [8] and SnSe. ...
... The synthesis uses Se powder and SnCl 2 · 2H 2 O as the precursors. [15,31,43] A comprehensive explanation of the chemistry of this synthesis requires balanced chemical equations and thermodynamic constants that describe the multiple equilibria. The accurate description of the chemical species present is complex, and far from the scope of this work. ...
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.
Article
Full-text available
Atomic-scale control of the chemical composition of semiconductor nanocrystals through a cation exchange reaction affords greater tunability in the design of multifunctional semiconductor composite nanocrystals. Here, we report a facile route to SnSe–Ag2Se composite nanocrystals using cation exchange at room temperature. Starting from freshly synthesized SnSe nanorods, we leverage the strong distortion of the Sn²⁺ octahedral coordination in SnSe and the hard–soft acid–base (HSAB) principle, to promote the exchange of an Sn²⁺ ion with two Ag⁺ ions in methanol leading to Ag2Se/SnSe nanocomposites. The morphology and chemistry of the nanocrystals evolve from nanorods with SnSe@Ag2Se (core@shell) structures for SnSe-rich composites to nanorods with a random distribution of Sn²⁺ and Ag⁺ ions for nearly equimolar composites, and finally to irregular fragmented nanocrystals for Ag2Se-rich composites. A mechanistic understanding of the observed morphology evolution is discussed using the change in the cation coordination from octahedral (Sn²⁺) to tetrahedral (Ag⁺) geometry and the accompanying expansion of the hcp Se²⁻ sublattice. Interestingly, the synthesized composite nanocrystals exhibit an optical band gap value tunable within a wide energy range by increasing the Ag2Se/SnSe ratio. This work provides a useful and facile strategy to modify the optical behavior of semiconductor nanomaterials, which can be leveraged for the design of better optical and/or photovoltaic devices.
Article
There is a growing interest in cost-effective polycrystalline SnSe-based thermoelectric (TE) materials, which are able to replace the high performance but mechanically fragile and costly single-crystalline SnSe. In this study, we present a low-temperature solution-based approach to produce SnSe-PbSe nanocomposites with outstanding TE performance. Our method involves combining surfactant-free SnSe particles with oleate-capped PbSe nanocrystals in specific ratios, followed by thermal annealing and consolidation using spark plasma sintering. These nanocomposites are characterized by distinct compositional and structural properties that significantly impact their transport properties. In particular, the addition of oleate-capped PbSe nanocrystals results in: i) a reduction in the electrostatically adsorbed Na at the surface of the SnSe particles; ii) a reduction of Sn vacancies due to alloying with Pb; iii) an increase in grain boundary density; and iv) the formation of PbSnSe secondary phases. Notably, the SnSe-2.5 %PbSe nanocomposites demonstrate a 30 % decrease in thermal conductivity compared to that of the SnSe matrix. This reduction contributes to a maximum figure of merit (zT) of 1.75 at 788 K with a high average zT value of ca. 1.2 in the medium temperature range of 573–773 K. These values represent one of the highest reported in polycrystalline SnSe materials, showcasing the potential of our fabricated SnSe-PbSe nanocomposites for cost-effective TE applications.
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
Downscaling of tin monoselenide (SnSe) samples to the nanometer regime (~80–20 nm) without affecting the structure, homogeneity, and optoelectronic properties was carried out by high-energy planetary ball milling (BM). The milling rate was varied from 200 rpm to 800 rpm by adopting a dry and wet-grinding top-down approach on customized stoichiometric SnSe precursors. The degree of crystallinity was assessed by powder x-ray diffraction (PXRD) and selected area electron diffraction. The lattice parameters, a = 4.435 Å, b = 11.498 Å, and c = 4.148 Å, of the nanoparticles were calculated from the PXRD data. Energy-dispersive x-ray analysis confirmed the chemical homogeneity (49.88:51.12 at.%) of the samples. The effects of rotational velocity as well as mode of grinding on the morphology and the size of SnSe powders were investigated using electron microscopes. The direct optical transition with band gap varied from 1.75 eV to 2.28 eV was elucidated from UV-Vis-NIR data. Photoluminescence revealed an increase in the intensity of the emission peak at 462.97 nm with angular velocities for both types of grinding. The variation of electrical resistivity (36–107 Ω cm) and mobility (3.45–1.12 cm2/Vs) with rotational speed was calculated for all the samples. The results obtained for the ball-milled nanoparticles pave the way towards the reduction of particle size, formation of stable morphology, and appreciable crystalline structure quality suitable for solar cell absorbers.
Article
Full-text available
The unique electronic and catalytic properties emerging from low symmetry anisotropic (1D and 2D) metal chalcogenides (MCs) have generated tremendous interest for use in next generation electronics, optoelectronics, electrochemical energy storage devices, and chemical sensing devices. Despite many proof-of-concept demonstrations so far, the full potential of anisotropic chalcogenides has yet to be investigated. This article provides a comprehensive overview of the recent progress made in the synthesis, mechanistic understanding, property modulation strategies, and applications of the anisotropic chalcogenides. It begins with an introduction to the basic crystal structures, and then the unique physical and chemical properties of 1D and 2D MCs. Controlled synthetic routes for anisotropic MC crystals are summarized with example advances in the solution-phase synthesis, vapor-phase synthesis, and exfoliation. Several important approaches to modulate dimensions, phases, compositions, defects, and heterostructures of anisotropic MCs are discussed. Recent significant advances in applications are highlighted for electronics, optoelectronic devices, catalysts, batteries, supercapacitors, sensing platforms, and thermoelectric devices. The article ends with prospects for future opportunities and challenges to be addressed in the academic research and practical engineering of anisotropic MCs.
Article
SnSe single crystals have gained great interest due to their excellent thermoelectric performance. However, polycrystalline SnSe is greatly desired due to facile processing, machinability, and scale-up application. Here, we report an outstanding high average ZT of 0.88 as well as a high peak ZT of 1.92 in solution-processed SnSe nanoplates. Nanosized boundaries formed by nanoplates and lattice strain created by lattice dislocations and stacking faults effectively scatter heat-carrying phonons, resulting in an ultralow lattice thermal conductivity of 0.19 W m-1 K-1 at 873 K. Ultraviolet photoelectron spectroscopy reveals that Ge and In incorporation produces an enhanced density of states in the electronic structure of SnSe, resulting in a large Seebeck coefficient. Ge and In codoping not only optimizes the Seebeck coefficient but also substantially increases the carrier concentration and electrical conductivity, helping to maintain a high power factor over a wide temperature range. Benefiting from an enhanced power factor and markedly reduced lattice thermal conductivity, high average ZT and peak ZT are achieved in Ge- and In-codoped SnSe nanoplates. This work achieves an ultrahigh average ZT of 0.88 in polycrystalline SnSe by adopting nontoxic element doping, potentially expanding its usefulness for various thermoelectric generator applications.
Article
Full-text available
Thermoelectric technology, harvesting electric power directly from heat, is a promising environmentally-friendly means of energy savings and power generation. The thermoelectric efficiency is determined by the device dimensionless figure of merit ZTdev, and optimizing this efficiency requires maximizing ZT values over a broad temperature range. Herein, we report a record high ZTdev ∼1.34, with ZT ranging from 0.7 to 2.0 at 300-773K, realized in hole doped SnSe crystals. The exceptional performance arises from the ultra-high power factor, which comes from a high electrical conductivity and a strongly enhanced Seebeck coefficient enabled by the contribution of multiple electronic valence bands present in SnSe. SnSe is a robust thermoelectric candidate for energy conversion applications in the low and moderate temperature range.
Article
Full-text available
Polycrystalline SnSe was synthesized by a melting-annealing-sintering process. X-ray diffraction reveals the sample possesses pure phase and strong orientation along [h00] direction. The degree of the orientations was estimated and the anisotropic thermoelectric properties are characterized. The polycrystalline sample shows a low electrical conductivity and a positive and large Seebeck coefficient. The low thermal conductivity is also observed in polycrystalline sample, but slightly higher than that of single crystal. The minimum value of thermal conductivity was measured as 0.3 W/m·K at 790 K. With the increase of the orientation factor, both electrical and thermal conductivities decrease, but the thermopowers are unchanged. As a consequence, the zT values remain unchanged in the polycrystalline samples despite the large variation in the degree of orientation.
Article
Full-text available
SnSe has been prepared by arc-melting, as mechanically robust pellets, consisting of highly oriented polycrystals. This material has been characterized by neutron powder diffraction (NPD), scanning electron microscopy, and transport measurements. A microscopic analysis from NPD data demonstrates a quite perfect stoichiometry SnSe0.98(2) and a fair amount of anharmonicity of the chemical bonds. The Seebeck coefficient reaches a record maximum value of 668 μV K−1 at 380 K; simultaneously, this highly oriented sample exhibits an extremely low thermal conductivity lower than 0.1 W m−1 K−1 around room temperature, which are two of the main ingredients of good thermoelectric materials. These excellent features exceed the reported values for this semiconducting compound in single crystalline form in the moderate-temperatures region and highlight its possibilities as a potential thermoelectric material.
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.
Article
Lead-free polycrystalline SnSe is a promising thermoelectric compound consisting of earth-abundant elements.However, poor electrical transport property for low intrinsic defect concentration (3×1017 cm-3) limits the usage of stoichiometric SnSe. In this work, Na2Se as acceptors was doped into SnSe compound in order to optimize the electrical transport properties, especially to increase the carrier concentration. As a result, the carrier concentrations increased and saturated at about 1.0×1019 cm-3 for Na-doped Na0.01Sn0.99Se at 300K. A maximum power factor of 0.48 mWm-1K-2 was obtained. And a maximum zT value of 0.75 was obtained for Na-doped Na0.01Sn0.99Se along the direction perpendicular to the sintering pressure at 823 K, which is 25% higher than that (0.6) of the undoped SnSe compound.
Article
SnSe has attracted some attention due to the recent report of record high thermoelectric figure-of-merit (ZT) of ∼2.6 at about 923 K in single crystals. Even though the ZT is high in single crystal SnSe, it was very low in its polycrystalline form due to the very low electrical conductivity (<103 S m−1) from 300 K to 700 K. In this work, we report studies on enhancement of electrical conductivity by Na doping to optimize the carrier concentration in polycrystalline SnSe prepared by melting and hot pressing. A room temperature carrier concentration of ∼2.7 × 1019 cm−3 was obtained in 2 atm% Na-doped SnSe samples with the highest power factor obtained in 1.5 atm% Na-doping. A peak ZT of ∼0.8 was achieved at 773 K along the hot pressing direction and the average ZT was improved.
Article
Excellent thermoelectric performance is obtained over a broad temperature range from 300 K to 800 K by doping single crystals of SnSe. The average value of the figure of merit ZT, of more than 1.17, is measured from 300 K to 800 K along the crystallographic b-axis of 3 at% Na-doped SnSe, with the maximum ZT reaching a value of 2 at 800 K. The room temperature value of the power factor for the same sample and in the same direction is 2.8 mW mK−2, which is an order of magnitude higher than that of the undoped crystal. Calculations show that Na doping lowers the Fermi level and increases the number of carrier pockets in SnSe, leading to a collaborative optimization of the Seebeck coefficient and the electrical conductivity. The resultant optimized carrier concentration and the increased number of carrier pockets near the Fermi level in Na-doped samples are believed to be the key factors behind the spectacular enhancement of the average ZT.
Article
SnSe, a ‘‘simple’’ and ‘‘old’’ binary compound composed of earth-abundant elements, has been reported to exhibit a high thermoelectric performance in single crystals, which stimulated recent interest in its polycrystalline counterparts. This work investigated the electrical and thermal transport properties of pristine and Na-doped SnSe1�xTex polycrystals prepared by mechanical alloying and spark plasma sintering. It is revealed that SnSe1�xTex solid solutions are formed when x ranges from 0 to 0.2. An energy barrier scattering mechanism is suitable for understanding the electrical conducting behaviour observed in the present SnSe polycrystalline materials, which may be associated with abundant defects at grain boundaries. The thermal conductivity was greatly reduced upon Te substitution due to alloy scattering of phonons as well explained by the Debye model. Due to the increased carrier concentration by Na-doping, thermoelectric figure of merit (ZT) was enhanced in the whole temperature range with a maximum value of 0.72 obtained at a relatively low temperature (773 K) for Sn0.99Na0.01Se0.84Te0.16.
Article
Iodine-doped n-type SnSe polycrystalline by melting and hot pressing is prepared. The prepared material is anisotropic with a peak ZT of ≈0.8 at about 773 K measured along the hot pressing direction. This is the first report on thermoelectric properties of n-type Sn chalcogenide alloys. With increasing content of iodine, the carrier concentration changed from 2.3 × 1017 cm−3 (p-type) to 5.0 × 1015 cm−3 (n-type) then to 2.0 × 1017 cm−3 (n-type). The decent ZT is mainly attributed to the intrinsically low thermal conductivity due to the high anharmonicity of the chemical bonds like those in p-type SnSe. By alloying with 10 at% SnS, even lower thermal conductivity and an enhanced Seebeck coefficient were achieved, leading to an increased ZT of ≈1.0 at about 773 K measured also along the hot pressing direction.