You are on page 1of 9

View Article Online

View Journal

Nanoscale
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: C. Wang, Z. Ma, J.
Lei, D. Zhang, Y. Chen, Y. Wang, J. Wang and Z. X. Cheng, Nanoscale, 2020, DOI: 10.1039/C9NR09331A.
Volume 10
Number 4
28 January 2018
This is an Accepted Manuscript, which has been through the
Pages 1549-2172
Royal Society of Chemistry peer review process and has been
accepted for publication.
Nanoscale Accepted Manuscripts are published online shortly after acceptance,
rsc.li/nanoscale

before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 2040-3372 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
Shuping Xu, Chongyang Liang et al.
Organelle-targeting surface-enhanced Raman scattering
shall the Royal Society of Chemistry be held responsible for any errors
(SERS) nanosensors for subcellular pH sensing

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/nanoscale
Page 1 of 8 Nanoscale
View Article Online
DOI: 10.1039/C9NR09331A

J
our
nalName

Core-shell nanostructures introduce multiple potential


barriers to enhance energy filtering for the improve-
ment of thermoelectric properties of SnTe
Zheng Ma,a Chao Wang,∗a Jingdan Lei,a De Zhang,a Yanqun Chen,a Yuanxu Wang,a
Jianli Wang,∗a and Zhenxiang Cheng∗ab

Nanoscale Accepted Manuscript


Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

Using dispersed nanostructures to induce energy filtering effect is an easy and effective mecha-
nism to optimize the performance of bulk thermoelectric materials. Comparing with other nanos-
tructures, core-shell nanostructures possess more interfaces and multiple potential barriers, which
would lead to significant impact on the thermal and electrical properties of materials. In this pa-
per, after BiCuSeO alloy doping into SnTe, SnO2 layers were formed at the interfaces and the Bi-
CuSeO nanoparticles would be wrapped in SnO2 shell during the following high temperature solid
state reaction. The formation of SnO2 layers could be observed and confirmed by x-ray diffrac-
tion (XRD) and scanning electron microscope (SEM). BiCuSeO@SnO2 core-shell nanostructures
can introduce multiple potential barriers to enhance energy filtering effect. Once the BiCuSeO
doping concentration were over 3%, the carrier concentration could decrease to about 10% while
the mobility increase to 350% comparing to the value of undoped sample at room temperature.
Meanwhile, the Seebeck coefficients were improved to 176.05 µ VK−1 at 835 K. Additionally, due
to the scattering of core-shell nanostructures for the phonons, the lower thermal conductivity is
achieved with the value of 1.04 Wm−1 K−1 at 835 K in Sn1.03 Te-5% BiCuSeO. Combining with the
improvement of thermal and electrical properties by the BiCuSeO@SnO2 core-shell, a high ZT
value ∼ 1.21 was achieved for the Sn1.03 Te-5% BiCuSeO at 835 K, which was enhanced 190%
than pristine SnTe.

1 Introduction lattice thermal conductivity, respectively. It is obvious that high


Despite a growing shortage of non-renewable energy, more than ZT materials must have a high Seebeck coefficient, high electri-
half of all energy is still wasted as heat 1,2 . Thermoelectric tech- cal conductivity and low thermal conductivity. However, these
nology is the simplest technology that can realize the direct con- parameters have strong coupling characteristics because they are
version of thermal energy and electric energy. It can convert solar determined by more fundamental characteristics of the electron
energy, geothermal energy, motor vehicles and industrial waste and phonon systems 3,4 . Increasing Seebeck coefficient usually
heat into electricity. However, the biggest challenge in the devel- results in a decrease of the electrical conductivity. Similarly, a de-
opment of thermoelectric technology is the improvement of ther- crease in thermal conductivity inevitably worsens electrical con-
moelectric performance of materials. The performance of a ther- ductivity. So far, nanostructures 5–7 , increase the slope of the elec-
moelectric material is related to the figure of merit ZT =S2 σ T/κ , tronic density of states 8,9 , energy filtering 10 and other strategies
where σ is the electrical conductivity, S the Seebeck coefficient, can effectively weaken the coupling relationship between various
κ the total thermal conductivity and T the absolute temperature, parameters 11,12 .
respectively. The total thermal conductivity, κ =κL +κe , consists of In the past decade, PbTe-based thermoelectric materials have
two components, namely κe and κL , which are the electronic and made remarkable progress with the optimum content of ZT ex-
ceeding 2, especially for p-type compositions 2,13,14 . However,
due to the toxicity of Pb, the popularization and application of
a
Institute for Computational Materials Science, School of Physics and Electronics, PbTe-based thermoelectric materials is limited. Another attractive
Henan University, Kaifeng, 475004, China. E-mail: wangchao@vip.henu.edu.cn, thermoelectric material compound, SnTe, has the similar crystal
jianli@uow.edu.au
b structure and electronic structure to PbTe. Therefore, SnTe is con-
Institute for Superconducting and Electronic Materials, University of Wollongong,
Squires Way, North Wollongong, Australia. E-mail: cheng@uow.edu.au sidered as one of the most promising alternatives to lead-free
thermoelectricity 15–19 , whereas the thermoelectric performance

J
our
nal
Name,
[yea
r][
,vol
.
], 1–7 | 1
Nanoscale Page 2 of 8
View Article Online
DOI: 10.1039/C9NR09331A

of SnTe is not as good as that of PbTe 20,21 . The poor intrinsic tronic thermal conductivity, thus further optimizing the thermo-
thermoelectric performance results from the very high hole con- electric properties of materials. Finally, a high Seebeck coefficient
centration (1020 ∼1021 cm−3 ) due to the intrinsic Sn vacancies, of 176.05 µ VK−1 and an extremely low thermal conductivity of
which led to low Seebeck coefficient and high electronic thermal 1.04 Wm−1 K−1 at 835 K can be obtained, which also results in a
conductivity 22,23 . In addition, although SnTe has a similar band very high ZT value of 1.21 at 835 K for Sn1.03 Te-5% BiCuSeO.
structure to PbTe, its light (L point) and heavy (Σ point) valence
band energy separation is larger (0.35 eV at 300 K) and the band 2 Experimental section
gap is smaller (0.18 eV at 300 K) 24,25 . This further leads to very In the preparation of materials, we set the ratio of Sn to Te as
small Seebeck coefficient for the pristine SnTe 26 . It can be seen 1.03:1 to solve the deterioration of material properties caused
that the most important way to optimize the thermoelectric per- by Sn volatilization in the preparation process. To synthesize
formance of SnTe is to improve Seebeck coefficient, as well as Sn1.03 Te, starting materials with suitable proportion, Sn (powder,
reducing thermal conductivity. Aladdin, 99.9%) and Te (powder, Aladdin, 99.9%), were mixed
The relation between Seebeck coefficient and electrical conduc- and ground in a planetary ball mill (MSK-SFM-1, Hefei Kejing
tivity can be expressed as 27 : Materials Technology Co. Ltd.) at 150 rpm for 4 h and placed

Nanoscale Accepted Manuscript


into a quartz tube. Then the quartz tube was evacuated to high
Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

{ }
π 2 kB d[ln(σ (E))] vacuum (∼2.5×10−2 Pa) and sintered at 1173 K for 15 h inside
S= kB T
3 q dE E=EF a muffle furnace (KSL-1200X, Hefei Kejing Materials Technology
{ } Co. Ltd.). After the quartz tube cooled to room temperature, the
π 2 kB 1 dn(E) 1 d µ (E) sintered material was put into a ball mill to crush into powder.
= kB T + , (1)
3 q n dE µ dE E=EF Finally, the crushed material was sintered in a 12.7 mm diameter
where n(E) and µ (E) are energy dependent carrier density and mold into a wafer by a spark plasma sintering system (SPS-211LX,
mobility, respectively. It can be seen from Equation 1 that there Fuji Electronic Industrial Co. Ltd.). The similar preparation pro-
are two mechanisms to enhance the Seebeck coefficient: (I) in- cess was used to BiCuSeO (wt %) alloy powder doped SnTe. The
creasing the density of states near the Fermi level, e.g., reso- production process for BiCuSeO was same as previous reports 36 .
nant level or valence band convergence 27–29 , and (II) increas- The BiCuSeO alloy particles range in size from a few nanome-
ing energy dependence of µ (E) by energy filtering (or carrier ters to a few microns and BiCuSeO nanoparticles only refer to the
filtering) 30 . L. W. Fu et al. showed that enhancement of ther- particles with size below 100 nm.
moelectric properties of Yb-filled skutterudites by an Ni-induced Crystal structure of the obtained samples were characterized
"core-shell" structure 31 ; C. L. Ou et al. demonstrated that the by X-ray diffraction using Bruker D8 Advance with Cu Kα ra-
introduction of double-barrier through TiC1−x Ox @TiOy -TiO2 het- diation accelerated in a 2θ range of 20-80◦ . The Synchrotron
erostructure improved the Seebeck coefficient of the material 32 , powder diffraction (SPD) patterns were collected on the Powder
and so on 33,34 . In this work, we introduced BiCuSeO@SnO2 Diffraction beamline at the Australian Synchrotron with a wave-
core-shell nanostructures into SnTe matrix by doping BiCuSeO length of 0.727464 Å. During the SPD measurement, the sam-
nano-alloy. During high-temperature solid state reaction, Sn ion ple was heated from 300 to 700 K. The microstructure examined
in SnTe matrix can react with O ion in BiCuSeO surface to form on a freshly broken surface of the samples was observed by a
SnO2 envelope, thus forming BiCuSeO@SnO2 core-shell nanos- field emission scanning electron microscope (FESEM, JSM-7001F,
tructures. BiCuSeO is composed by alternately stacking the in- JEOL Co., Ltd.). The obtained pellets after spark plasma sintering
sulating oxide (Bi2 O2 )2+ layer acting as charge reservoir and the processed were cut into bars with dimensions of 12 mm × 2 mm
conductive selenide (Cu2 Se2 )2− layer constituting a conduction × 3 mm that were used to measure electrical conductivity and
pathway for carrier transport 35,36 . The presence of layered struc- Seebeck coefficient simultaneously using a static DC thermoelec-
ture leads to the formation of more complex nonplanar poten- tric property measurement system (ZEM-3, ULVAC-RIKO, Inc.)
tial barrier. At the same time, as a shell, SnO2 can form energy under a low-pressure He atmosphere from room temperature to
barriers on the interface of contact with SnTe matrix and core 835 K. Carrier concentration and mobility were analyzed at room
due to its wide bandgap (Eg = 3.59eV 37 ). Therefore, the pres- temperature using DC Hall measurement system (ET9005, East
ence of BiCuSeO@SnO2 core-shell nanostructures can introduce Changing Technologies, Inc.). The thermal conductivity was de-
multiple potential barriers in the matrix to induce an enhanced termined from the thermal diffusivity obtained by the laser flash
energy filtering effect. The strengthened energy filtering effect method (DLF-1/EM1200, TA Inc.) in a Ar atmosphere. In our
can more effectively prevent low-energy carriers from participat- work, the density was determined using the dimensions and mass
ing in conductivity through the material. Higher energy electrons of the sample. A typical disk shaped sample is obtained which is
can cross these energy barriers. Because the high-energy carriers 12.7 mm in diameter with density no less than 96% of theoretical
are nothing but the matter of energy transfer, the Seebeck coeffi- density (6.46 gcm−3 ).
cient gradually increase. It is worth mentioning that this ability
is not possessed by sole energy barriers. The ineffectiveness of
3 Results and Discussion
the sole energy barriers to enhance thermoelectric efficiency has Fig. 1a) shows the XRD patterns of the Sn1.03 Te-x% BiCuSeO (x =
been rather extensively proved in the literature 38,39 . The exis- 0, 1, 2, 3, 4, 5 and 6) samples. The position of the Bragg diffrac-
tence of enhanced energy filtering effect also inhibits the elec- tion peak exhibit a phase that can be indexed to the rock-salt SnTe

2| J
our
nal
Name,
[yea
r][
,vol
.
],
1–7
Page 3 of 8 Nanoscale
View Article Online
DOI: 10.1039/C9NR09331A

a) SnO2 SnO2 SnO2


b) SnO 2
SnO2
SnO2
x=6
700K
x=5
x=4
x=3 600K
x=2
x=1
x=0 500K
BiCuSeO
400K
300K

c) Yobs
Intensity (arb.units)

Ycalc
Yobs-Ycalc
Bragg-position

Nanoscale Accepted Manuscript


Fig. 2 FESEM images: a) Sn1.03 Te sample, b)-d) doped 1%, 2%, 3%
Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

BiCuSeO alloy, respectively; g)-h) doped 4%,5% BiCuSeO alloy,


4 7 10 13 16 19 22 25 28 31 34 37 respectively ; i) and j) doped 6% BiCuSeO alloy. e) and f) The energy
2θ dispersive spectrum of 3% and 4% BiCuSeO alloy , respectively. d1)-d6)
The elemental energy dispersive spectroscopy (EDS) mapping of the
Fig. 1 a) XRD patterns for Sn1.03 Te-x% BiCuSeO samples and doped Sn1.03 Te-3% BiCuSeO sample.
BiCuSeO alloy powder; b) Synchrotron powder diffraction (SPD)
patterns of Sn1.03 Te-5% BiCuSeO collected with the temperature
increasing from 300 to 700 K. c) (Color online) Results of the Rietveld
profile refinement for the low-temperature face-centered cubic phase of In Fig. 2a)-d) and g)-j), we can clearly see that the BiCuSeO alloy
Sn1.03 Te-5% BiCuSeO. The green small vertical lines indicate the particles were embedded into SnTe matrix and the size of alloy
positions of the Bragg peaks corresponding to the Sn1.03 Te phases and
particles ranges from tens of nanometers to several micrometers.
SnO2 phases. The blue curves at the bottom denote differences
between the measured and calculated intensities. EDS images are shown in Fig. 2e) and f), which indicated obvious
Bi, Cu, Se and O peaks. The elements EDS mapping for the sam-
ple Sn1.03 Te-3% BiCuSeO in Fig. 2 d1)-d6) exhibit Sn, Te, Bi, Cu,
(Fm3̄m) structure. In addition, the diffraction peak of the phase Se, and O distribution, respectively. Some dark areas appear in
of BiCuSeO was not detected by X-ray characterization, but the EDS mapping of Te, which are corresponding to the bright areas
diffraction peak of the second phase SnO2 was found. The exis- in EDS mapping of Cu and O. Therefore, the presence of BiCuSeO
tence of a large amount of SnO2 second phase is caused by the alloy particles was further confirmed. Especially, these areas are
reaction of O ions in BiCuSeO with Sn ions at high-temperature also bright in EDS mapping of Sn. Combining with the XRD pat-
solid state. For Sn to react with O ions in BiCuSeO, it is essentially terns, we can conclude that the surface of BiCuSeO alloy particles
to react with O ions in (Bi2 O2 )2+ layer. Sn is less electronegative was covered by SnO2 . Thus, the existence of BiCuSeO@SnO2
than Bi, that is, the electronegative of Sn and Bi are 1.477 and core-shell structures are proved.
2.342 respectively 40 . In other words, Bi element has a stronger TEM images of the BiCuSeO@SnO2 core-shell nanostructures
oxidation ability than Sn element. Moreover, SnO2 ’s Gibbs free after SPS treatment using the obtained Sn1.03 Te-5% BiCuSeO
energy is smaller than that of Bi2 O3 , so SnO2 is easier to gen- powder composites are presented in Fig. 3. The region I (the
erated than Bi2 O3 41 . Therefore, the BiCuSeO nanoparticles dis- area enclosed by the yellow dotted line) in Fig. 3 a)-c) is the
tributed in SnTe matrix are coated with a layer of SnO2 to from core structure formed by BiCuSeO nanoparticles, and the region
BiCuSeO@SnO2 core-shell nanostructures. In order to more ac- II (the area between the dotted yellow line and the solid red line)
curately explore the crystal structure of the sample, Sn1.03 Te-5% is the shell structure formed by SnO2 . EDS pattern analysis in
BiCuSeO sample was studied using the SPD with the temperature Fig. 3 d) also proves that region II is SnO2 . In Fig. 3 c), we can
increasing from 300 to 700 K, as shown in Fig. 1b). In addition see that the outer layer of BiCuSeO nanoparticles as the nucleus
to the results similar to those of X-ray diffraction, the SPD exper- is covered by an insulating layer SnO2 that is different from the
iment also proved that the sample structure did not change with core material. This result is similar to other experimental reports
the increase of temperature. This means that BiCuSeO@SnO2 on core-shell structures 42,43 . Thus the existence of core-shell
core-shell nanostructures is always present in the test temper- structure is further confirmed. The high magnification in Fig. 3
ature range. Results of the Rietveld refinements performed at a) and b) reveals that the dispersed BiCuSeO@SnO2 core-shell
low-temperature structure is shown in Fig. 1c). The fits are ex- nanostructures are in incoherent phase with the pure SnTe ma-
ceptionally good, which vouches for the high quality of our data trix, and the interface boundary between the matrix phase SnTe
and lends confidence to the claim that the cubic model accurately and the BiCuSeO@SnO2 core-shell nanostructures can be clearly
describes the average and local structure at this temperature. observed. Such formations of interfaces act as multiple potential
A scanning electron microscopy (SEM) instrument equipped barriers which provide the energy filtering behavior.
with an energy dispersive spectrometer (EDS) apparatus was used Further understanding of energy filtering mechanism, the hall
to further characterize the phase composition, as shown in Fig. 2. test on Sn1.03 Te-x% BiCuSeO samples were carried out, and

J
our
nal
Name,
[yea
r][
,vol
.
], 1–7 | 3
Nanoscale Page 4 of 8
View Article Online
DOI: 10.1039/C9NR09331A

shown in Fig. 4a). It is noted that carrier concentration grad-


ually decreased with the increase of BiCuSeO doped concentra-
tion. Once the BiCuSeO doping concentration were over 3%,
the carrier concentration could decrease to about 10% while the
mobility increase to 350% comparing to the value of undoped
sample at room temperature. This phenomenon is caused by a
band-structure mismatch between the dispersed BiCuSeO@SnO2
core-shell nanostructures and the host material 47,48 , that is be-
cause of large band gap and less number electrons for trans-
portation about SnO2 layer. In other words, the formation of the
BiCuSeO@SnO2 core-shell nanostructures results in a heteroge-
neous interface between the interface of the matrix and the shell,
the interface between the shell and the core, and the layers within
the core. The heterogeneous interfaces can act as an energy bar-

Nanoscale Accepted Manuscript


rier to substantially scatter low-energy carriers; thus, high-energy
Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

carriers can tunnel through the thin SnO2 layers, while the low-
energy carries can not tunnel through. In addition, the hall mobil-
ity was enhanced by the dispersion of BiCuSeO@SnO2 core-shell
Fig. 3 a)-b) High-magnification TEM image of the Sn1.03 Te-5% nanostructures, and it is expected that the formation of asym-
BiCuSeO sample; c) STEM image of the Sn1.03 Te-5% BiCuSeO sample;
d) EDS pattern of region II (area II refers to the area between the dotted
metric potentials from the BiCuSeO@SnO2 core-shell nanostruc-
yellow line and the solid red line). tures within the host matrix could weaken scattering of the ma-
jority carriers 48 . It is worth noting that the introduction of Bi-
CuSeO nano-alloy can significantly improve hall mobility and ob-
viously reduce carrier concentration compared with other doped
a) b) nano-alloys 18,49–51 . This is due to the fact that the doping of
BiCuSeO@SnO2 core-shell nanostructures can introduce multiple
potential barriers in the matrix to induce an enhanced energy fil-
tering effect. In a nutshell, energy filtering occurs when a poten-
tial barrier in an otherwise uniform medium prevents the diffu-
sion/drift of carriers with energy E lower than the barrier height
Vb 38 . The relationship between potential barriers and thermo-
electric properties has been studied in theory. Use of potential
Fig. 4 a) The carrier concentration and mobility of Sn1.03 Te-x%
BiCuSeO at room temperature; b) Electrical conductivities as a function barriers to increase ZT through energy filtering was advanced by
of temperature for Sn1.03 Te-x% BiCuSeO. Nishio and Hirano, who estimated the optimal barrier height and
spacing using the Boltzmann transport equation (BTE) 52 . Ka-
jikawa applied the energy filtering model in BTE to the polycrys-
talline semiconductor, and obtained the analytic expressions of S
Pristine SnTe grain boundary Dopde SnTe and σ , assuming energy barriers of uniform heights. 53,54 . Z. X.
BiCuSeO@SnO
et al. then obtained a simple generic parametric equations are
core-shell nanostructures found that are in agreement with the exact Boltzmann transport
I Core : BiCuSeO
formalism in a wide range of parameters 55 . For a bulk semi-
conductor and for Vb − EF ≥ 2kB T, simple generic parametric
II Shell : SnO equations analytical expressions for the conductivity σ and the
Seebeck coefficient S can be written as
CB SnTe SnO2 BiCuSeO SnO2 SnTe
+
h (low E)
CB CB CB CB CB σ = σC Fp (xb ) (2)
EF Eg=1eV EF
+
h (high E) Eg=0.18eV
VB Eg=3.59eV
VB
and [ ( ) ]
VB VB VB VB kB 5
No Energy Filtering
S=− ηF − p + − ∆p (xb ) (3)
e 2
Fig. 5 A schematic diagram of doped BiCuSeO@SnO2 core-shell where ( )
4e2 p+ 23 3 ηF 5
nanostructures introducing multiple potential barriers. The color σC = N0 τ 0 (kB T ) e Γ p + (4)
illustration is present band diagrams model for BiCuSeO@SnO2 3m∗ 2 2
core-shell nanostructures heterostructured interface. The work functions
with
for SnTe, SnO2 and BiCuSeO are 5.1 eV, 4.34 eV and 3.94 eV, ∞ xk
p+ 25
respectively 44–46 . Fp (xb ) = 1 − e−xb xb ∑ Γ (k + bp + 7 ) (5)
k=0 2

4| J
our
nal
Name,
[yea
r][
,vol
.
],
1–7
Page 5 of 8 Nanoscale
View Article Online
DOI: 10.1039/C9NR09331A

a) b) a) b)

Fig. 6 a) Seebeck coefficients as a function of temperature for c)


Sn1.03 Te-x% BiCuSeO; b) Room temperature Seebeck coefficient
versus carrier concentration (the solid line is Pisarenko curve based on
two-valence-band model). In the calculation, Eg =0.18 eV; ∆E=0.35 eV;
mlight =0.168; mheavy =1.92.

and

Nanoscale Accepted Manuscript


p+ 5
e−xb xb 2
Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

Fig. 7 a) Total thermal conductivity, b) lattice thermal conductivity and c)


∆p (xb ) = ( ) (6) electronic thermal conductivity as functions of temperature for
Γ p + 52 Fp (xb ) Sn1.03 Te-x% BiCuSeO.
In the above equations, Γ(z) is the gamma function, EF is the
Fermi level, and ηF =EF /kB T, xb = Vb /kB T are the reduced Fermi
energy and barrier height, respectively. The parameter p charac- coefficient as a function of the hole concentration, the so-called
terizes the type of scattering, being p = -1/2 for phonon scatter- Pisarenko curve, is presented in Fig. 6b), which is calculated ac-
ing and for highly screened ionized impurity scattering, and p = cording to two-valence-band model. It can be seen that the See-
3/2 for ionized impurity scattering 56 . The parameter k can sim- beck coefficient of the doped sample is above the dark solid curve.
ply be thought of as the number of barriers 55 . At the same time, The Seebeck coefficient values of nanocomposites must lie on the
it should be clear that the barrier height introduced by the core- Pisarenko line (dark solid curve) if there is no energy filtering
shell nanostructures is not all the same, but this simple generic effect or band engineering effect 57 . As is known to all, doping
parametric equations still applies. Therefore, it can be seen that Bi,Cu and Se elements cannot introduce band engineering effect
the conductivity of the material decreases and the Seebeck co- (valence band convergence) in SnTe matrix 16,26,58 . Therefore,
efficient increases as the height and number of potential barri- the Seebeck coefficient of the Sn1.03 Te-x% BiCuSeO samples at
ers increase. In the SnTe-BiCuSeO@SnO2 heterogeneous inter- room temperature occur above the dark solid curve, strengthen-
faces, a mass of low energy carriers can be filtered by multiple ing the notion of an enhanced energy filtering effect.
potential barriers, thus the electrical conductivity significantly de- The heterostructured interfaces act as scattering centers against
creases. From Fig. 4b), we can clearly see that the electrical con- phonons. They can effectively scatter middle and low frequency
ductivity significantly decreases in the whole temperature range phonons 59 . Therefore, the thermal conductivity of the sample de-
due to the introduction of BiCuSeO. For example, the electrical creases with the increase of doped concentration in the medium-
conductivity of Sn1.03 Te-6% BiCuSeO at room temperature was low temperature region, as shown in Fig. 7a). It can be clearly
2.73×103 Scm−1 , which was 55% lower than that of the pure seen from Fig. 7a) that the total thermal conductivity of Sn1.03 Te-
phase SnTe. Present band diagrams model for BiCuSeO@SnO2 x% BiCuSeO significantly decreases in the medium-low tempera-
core-shell nanostructures heterostructured interface is shown in ture region. The effect of reducing thermal conductivity due to
Fig. 5. The BiCuSeO@SnO2 core-shell nanostructures can build energy filtering effect is weak at high temperature. In order to
the multiple potential barriers filter the charge carrier transfer, further explore the reasons for the special dependence between
where the first interfacial SnO2 barrier at the SnTe-SnO2 interface Sn1.03 Te-x% BiCuSeO thermal conductivity and temperature, we
selectively scatters low-energy carriers rather than high-energy calculated the lattice thermal conductivity κL and the electronic
carriers, and the barrier at the BiCuSeO-SnO2 interface can fur- thermal conductivity κe and the results are shown in Fig. 7b) and
ther scatter partial high-energy carriers rather than higher ones. c). The electronic thermal conductivity κe can be calculated base
In addition, the barrier inside BiCuSeO also can re-filter the car- on the formula κe = Lσ T, where L is the Lorenz number. κL is
riers that have passed through the barrier at the SnO2 shell. Be- obtained using the equation κL =κ −κe . Here, the Lorenz num-
cause the high-energy carriers are nothing but the matter of en- ber is obtained based on the two band model and the calculated
ergy transfer, the Seebeck coefficient gradually increase. details is given in Supporting Information 60–66 . The variation of
As shown in Fig. 6a), the Seebeck coefficient of Sn1.03 Te-x% Lorenz number with the doping concentration is shown in Fig.
BiCuSeO samples gradually increases with the increase of tem- S1. We can find that κe of the sample significantly change with
perature. It can also be seen from the Fig. 6a) that the Seebeck the increase of the BiCuSeO doped concentration. As the tem-
coefficient of the sample doped with BiCuSeO is greater than that perature increases, more carriers would cross the energy barrier
of the pure phase SnTe. From Equation 1 and 3, we can also know due to the increase of the average energy of carriers, so the effect
that the energy filtering effect will effectively improve the See- of energy filtering on the thermal conductivity decrease. It can
beck coefficient of the material. The room-temperature Seebeck be seen from Fig. 7b) that the κL gradually decreases with the

J
our
nal
Name,
[yea
r][
,vol
.
], 1–7 | 5
Nanoscale Page 6 of 8
View Article Online
DOI: 10.1039/C9NR09331A

a) b) tures act as scattering centers against phonons. Therefore, the


lower thermal conductivity is achieved with the value of 1.04

0%
Wm−1 K−1 at 835 K in Sn1.03 Te-5% BiCuSeO. Finally, a high ZT

19
value ∼ 1.21 is achieved for the Sn1.03 Te-5% BiCuSeO at 835 K.

Acknowledgments
This research was sponsored by the National Natural Science
Fig. 8 a) Power factors and b) ZT as functions of temperature for of China (No. U1504511, 51371076, 51571083, 11674083),
Sn1.03 Te-x% BiCuSeO. the Science and Technology fund of Henan Province (No.
182102210227). We would like to thank the help from the Aus-
tralian Nuclear Science and Technology Organization and Aus-
increase of doping concentration, which is because the scatter- tralian Synchrotron during our experiments.
ing probability of medium-low frequency phonons increases with
the increase of the number of BiCuSeO nanoparticles. However, References

Nanoscale Accepted Manuscript


Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

with the increase of temperature, high-frequency phonons heat


1 K. Nielsch, J. Bachmann, J. Kimling and H. Böttner, Adv. En-
transfer dominates, and the phonon scattering effect caused by
ergy Mater., 2011, 1, 713–731.
nanoparticles weakens 67,68 . Thus, the effect of doped BiCuSeO
on reducing the thermal conductivity is deteriorative at high tem- 2 K. Biswas, J. He, I. D. Blum, C. I. Wu, T. P. Hogan, D. N.
perature. Finally, the lower thermal conductivity is achieved with Seidman, V. P. Dravid and M. G. Kanatzidis, Nature, 2012,
the value of 1.04 Wm−1 K−1 at 835 K in Sn1.03 Te-5% BiCuSeO. 489, 414–418.
Fig. 8a) shows dependence of the power factor on temperature 3 M. Dresselhaus, G. Chen, Z. Ren, J.-P. Fleurial, P. Gogna, M. Y.
of Sn1.03 Te-x% BiCuSeO samples. It can be seen that the power Tang, D. Vashaee, H. Lee, X. Wang, G. Joshi and et al., MRS
factor of Sn1.03 Te-x% BiCuSeO is deteriorated in the low tempera- Proceedings, 2007, 1044, 1044U02–04.
ture region and strengthened in the high temperature region. The 4 Y. Wang, L. Yang, X. L. Shi, X. Shi, L. Chen, M. S. Dargusch,
essential reason of this phenomenon lies in the intrinsic relation J. Zou and Z. G. Chen, Adv. Mater., 2019, 1807916.
between Seebeck coefficient and conductivity. An overall increase 5 R. Venkatasubramanian, E. Silvola, T. Colpitts and
of the PF is then possible if the Seebeck coefficient enhancement B. O’Quinn, Nature, 2001, 413, 597–602.
compensates for the decrease of the conductivity. This is an in- 6 A. Balandin and K. L. Wang, J. Appl. Phys., 1998, 84, 6149–
evitable disadvantage when the core-shell structure is introduced 6153.
to optimize the thermoelectric properties of materials 69 . In addi- 7 J. Lei, D. Zhang, W. Guan, Z. Ma, Z. Cheng, C. Wang and
tion, due to the influence of strong energy filtering effect, the elec- Y. Wang, Appl. Phys. Lett., 2018, 113, 083901.
trical conductivity decreases significantly, which make it reason- 8 L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B, 1993, 47,
able that the power factor increases at high temperature. A high 12727–12731.
power factor of 15.17 µ W cm−1 K−2 is achieved for the Sn1.03 Te- 9 Y. Tian, M. R. Sakr, J. M. Kinder, D. Liang, M. J. Macdonald,
5% BiCuSeO at 835 K. Finally, high ZT value ∼ 1.21 is achieved R. L. Qiu, H. J. Gao and X. P. Gao, Nano Lett., 2012, 12, 6492.
for the Sn1.03 Te-5% BiCuSeO at 835 K, which is enhanced 190% 10 S. V. Faleev and F. m. c. Léonard, Phys. Rev. B, 2008, 77,
than pristine SnTe. 214304.
4 Conclusion 11 M. Hong, T. C. Chasapis, Z. Chen, L. Yang, M. G. Kanatzidis,
G. J. Snyder and J. Zou, ACS Nano, 2016, 10, 4719–4727.
In summary, the BiCuSeO@SnO2 core-shell nanostructures lead
12 M. Hong, J. Zou and Z. G. Chen, Adv. Mater., 2019, 31,
to a large number of heterogeneous interfaces, thus can intro-
1807071.
duce multiple potential barriers to enhance energy filtering ef-
fect. The enhanced energy filtering effect can significantly im- 13 G. Tan, F. Shi, S. Hao, L. D. Zhao, H. Chi, X. Zhang, C. Uher,
prove hall mobility and obviously reduce carrier concentration C. Wolverton, V. P. Dravid and M. G. Kanatzidis, Nat. Com-
comparing to the value of undoped sample at room temperature. mun., 2016, 7, 12167.
In other word, the strengthened energy filtering effect can more 14 T. Fu, X. Yue, H. Wu, C. Fu, T. Zhu, X. Liu, L. Hu, P. Ying,
effectively prevent low-energy carriers from participating in con- J. He and X. Zhao, J. Materiomics, 2016, 2, 141–149.
ductivity through the material. Because the high-energy carriers 15 G. Tan, L. D. Zhao, F. Shi, J. W. Doak, S. H. Lo, H. Sun,
are nothing but the matter of energy transfer, the Seebeck coef- C. Wolverton, V. P. Dravid, C. Uher and M. G. Kanatzidis, J.
ficient gradually increase. For example, a Seebeck coefficient of Am. Chem. Soc., 2014, 136, 7006–7017.
176.05 µ VK−1 and a high power factor of 15.17 µ W cm−1 K−2 are 16 G. Tan, F. Shi, J. Doak, H. Sun, L. D. Zhao, P. Wang, C. Uher,
achieved for the Sn1.03 Te-5% BiCuSeO at 835 K. In terms of op- C. Wolverton, V. Dravid and M. Kanatzidis, Energy Environ.
timizing thermal performance, BiCuSeO@SnO2 core-shell nanos- Sci., 2016, 8, 267–277.
tructures can not only reduce the electronic thermal conductivity 17 G. Tan, F. Shi, S. Hao, H. Chi, T. P. Bailey, L. D. Zhao, C. Uher,
by introducing enhanced energy filtering effect, but also reduce C. Wolverton, V. P. Dravid and M. G. Kanatzidis, J. Am. Chem.
the lattice thermal conductivity due to the core-shell nanostruc- Soc., 2016, 46, 11507.

6| J
our
nal
Name,
[yea
r][
,vol
.
],
1–7
Page 7 of 8 Nanoscale
View Article Online
DOI: 10.1039/C9NR09331A

18 A. Banik, U. S. Shenoy, S. Saha, U. V. Waghmare and press, 2014.


K. Biswas, J. Am. Chem. Soc., 2016, 138, 08382. 42 G. Zheng, S. de Marchi, V. López-Puente, K. Sentosun,
19 R. Moshwan, L. Yang, J. Zou and Z. G. Chen, Adv. Funct. L. Polavarapu, I. Pérez-Juste, E. H. Hill, S. Bals, L. M. Liz-
Mater., 2017, 27, 1703278. Marzán, I. Pastoriza-Santos et al., Small, 2016, 12, 3935–
20 A. Banik, U. S. Shenoy, S. Anand, U. V. Waghmare and 3943.
K. Biswas, Chem. Mater., 2015, 27, 581–587. 43 X. Qiao, B. Su, C. Liu, Q. Song, D. Luo, G. Mo and T. Wang,
21 H. Wu, C. Chang, D. Feng, Y. Xiao, X. Zhang, Y. Pei, L. Zheng, Adv. Mater., 2018, 30, 1702275.
D. Wu, S. Gong and Y. Chen, Energy Environ. Sci., 2015, 8, 44 L. Chang, P. Stiles and L. Esaki, J. Appl. Phys., 1967, 38, 4440–
3298–3312. 4445.
22 M. J. Wahila, Z. W. Lebens-Higgins, N. F. Quackenbush, 45 M. Islam and M. Hakim, J. Mater. Sci. Lett., 1986, 5, 63–65.
J. Nishitani, W. Walukiewicz, P. A. Glans, J. H. Guo, J. C. 46 A. Achour, J. Liu, P. Peng, C. Shaw and Z. Huang, ACS Catal.,
Woicik, K. M. Yu and L. F. J. Piper, Phys. Rev. B, 2015, 91, 2018, 8, 10164–10172.
205307. 47 Y. Li, D. Li, X. Qin, X. Yang, Y. Liu, J. Zhang, Y. Dou, C. Song
23 J. He, X. Tan, J. Xu, G. Q. Liu, H. Shao, Y. Fu, W. Xue, L. Zhu, and H. Xin, J. Mater. Chem. C, 2015, 3, 7045–7052.

Nanoscale Accepted Manuscript


Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.

J. Xu and H. Jiang, J. Mater. Chem. A, 2015, 3, 19974–19979. 48 Y. Li, X. Qin, D. Li, J. Zhang, C. Li, Y. Liu, C. Song, H. Xin and
24 J. A. Kafalas, R. F. Brebrick and A. J. Strauss, Appl. Phys. Lett., H. Guo, Appl. Phys. Lett., 2016, 108, 062104.
1964, 4, 93–94. 49 G. Tan, S. Hao, R. C. Hanus, X. Zhang, S. Anand, T. P. Bailey,
25 L. M. Rogers, J. Phys. D: Appl. Phys., 1968, 1, 1067–1070. A. J. Rettie, X. Su, C. Uher, V. P. Dravid et al., ACS Energy Lett.,
26 R. Brebrick, J. Phys. Chem. Solids, 1963, 24, 27–36. 2018, 3, 705–712.
27 J. P. Heremans, J. Vladimir, E. S. Toberer, S. Ali, K. Ken, 50 Z. Ma, C. Wang, J. Lei, D. Zhang, Y. Chen, J. Wang, Z. Cheng
C. Anek, Y. Shinsuke and S. G Jeffrey, Science, 2008, 321, and Y. Wang, ACS Appl. Energy Mater., 2019, 2, 7354–7363.
554–557. 51 G. Tan, F. Shi, H. Sun, L.-D. Zhao, C. Uher, V. P. Dravid and
28 Y. Pei, X. Shi, A. Lalonde, H. Wang, L. Chen and G. J. Snyder, M. G. Kanatzidis, J. Mater. Chem. A, 2014, 2, 20849–20854.
Nature, 2011, 473, 66. 52 Y. Nishio and T. Hirano, Jpn. J. Appl. Phys., 1997, 36, 170.
29 Z. Ma, J. Lei, D. Zhang, C. Wang, J. Wang, Z. Cheng and 53 Y. Kajikawa, J. Appl. Polym. Sci., 2012, 112, 123713.
Y. Wang, ACS Appl. Mater. Interfaces, 2019, 11, 33792–33802. 54 Y. Kajikawa, J. Appl. Phys., 2013, 114, 053707.
30 W. Liu, X. Tan, K. Yin, H. Liu, X. Tang, J. Shi, Q. Zhang and 55 X. Zianni and D. Narducci, J. Appl. Phys., 2015, 117, 035102.
C. Uher, Phys. Rev. Lett., 2012, 108, 166601. 56 O. Jonasson and I. Knezevic, Phys. Rev. B, 2014, 90, 165415.
31 L. Fu, J. Yang, J. Peng, Q. Jiang, Y. Xiao, Y. Luo, D. Zhang, 57 T. Zou, X. Qin, Y. Zhang, X. Li, Z. Zeng, D. Li, J. Zhang, H. Xin,
Z. Zhou, M. Zhang, Y. Cheng and F. Cheng, J. Mater. Chem. A, W. Xie and A. Weidenkaff, Sci. Rep., 2015, 5, 17803.
2015, 3, year. 58 A. Banik and K. Biswas, J. Mater. Chem. A, 2014, 2, 9620–
32 C. Ou, J. Hou, T.-R. Wei, B. Jiang, S. Jiao, J. Li and H. Zhu, 9625.
NPG Asia Mater., 2015, 7, e182. 59 Y. Pei, J. Lensch-Falk, E. S. Toberer, D. L. Medlin and G. J.
33 M. Scheele, N. Oeschler, I. Veremchuk, S.-O. Peters, A. Littig, Snyder, Adv. Funct. Mater., 2011, 21, 241–249.
A. Kornowski, C. Klinke and H. Weller, ACS Nano, 2011, 5, 60 M. Zhou, Z. M. Gibbs, H. Wang, Y. Han, C. Xin, L. Li and G. J.
8541–8551. Snyder, Phys. Chem. Chem. Phys., 2014, 16, 20741–20748.
34 B. Madavali, H. Kim, K. Lee and S. Hong, Intermetallics, 2017, 61 Y. W. Tung and M. L. Cohen, Phys. Rev., 1969, 180, 823–826.
82, 68–75. 62 S. Rabii, Phys. Rev., 1969, 182, 821–828.
35 Y. Liu, L.-D. Zhao, Y. Liu, J. Lan, W. Xu, F. Li, B.-P. Zhang, 63 T. Seddon, S. C. Gupta and G. A. Saunders, Solid State Com-
D. Berardan, N. Dragoe, Y.-H. Lin, C.-W. Nan, J.-F. Li and mun., 1976, 20, 69–72.
H. Zhu, J. Am. Chem. Soc., 2011, 133, 20112–20115.
64 Y. I. Ravich, B. A. Efimova and V. I. Tamarchenko, Phys. Status
36 J. Lei, W. Guan, D. Zhang, Z. Ma, X. Yang, C. Wang and Solidi B, 2010, 4, A4–A8.
Y. Wang, Appl. Surf. Sci., 2019, 473, 985–991.
65 L. M. Rogers, J. Phys. D: Appl. Phys., 1968, 1, 1067.
37 C. Authors and editors of the volumes III/17E-17F-41C, Non-
66 Q. Zhang, F. Cao, W. Liu, K. Lukas, B. Yu, S. Chen, C. Opeil,
Tetrahedrally Bonded Elements and Binary Compounds I, 1998,
D. Broido, G. Chen and Z. Ren, J. Am. Chem. Soc., 2012, 134,
1–9.
10031.
38 D. Narducci, S. Frabboni and X. Zianni, J. Mater. Chem. C,
67 Z. Chen, Z. Jian, W. Li, Y. Chang, B. Ge, R. Hanus, J. Yang,
2015, 3, 12176–12185.
Y. Chen, M. Huang and G. J. Snyder, Adv. Mater., 2017, 29,
39 M. Bachmann, M. Czerner and C. Heiliger, Phys. Rev. B, 2012, 1606768.
86, 115320.
68 Z. Chen, B. Ge, W. Li, S. Lin, J. Shen, Y. Chang, R. Hanus,
40 R. T. Sanderson, J. Am. Chem. Soc., 1983, 105, 2259–2261. G. J. Snyder and Y. Pei, Nat. Commun., 2017, 8, 13828.
41 W. M. Haynes, CRC handbook of chemistry and physics, CRC 69 C. Gayner and Y. Amouyal, Adv. Funct. Mater., 2019, 1901789.

J
our
nal
Name,
[yea
r][
,vol
.
], 1–7 | 7
Published on 20 December 2019. Downloaded on 1/2/2020 11:04:46 PM.
Nanoscale

DOI: 10.1039/C9NR09331A
View Article Online

Nanoscale Accepted Manuscript


Page 8 of 8

You might also like