Aenm 201700446
Aenm 201700446
Aenm 201700446
Huaizhou Zhao1*, Binglei Cao1, 2, Shanming Li1, Ning Liu1, Jiawen Shen3, Shan Li1, Jikang
Jian2, Lin Gu1, Yanzhong Pei3, G. Jeffrey Snyder4, Zhifeng Ren5 and Xiaolong Chen1
1
Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of
3
School of Materials Science and Engineering, Tongji University, Shanghai 201804, China
4
Department of Materials Science and Engineering, Northwestern University, Evanston, Illinois 60208,
USA.
5
Department of Physics and TcSUH, University of Houston, Houston, TX 77204, USA
This is the author manuscript accepted for publication and has undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may lead to
differences between this version and the Version of Record. Please cite this article as doi:
10.1002/aenm.201700446.
1. Introduction
Recent studies have witnessed the fast growth of TE materials. Among the notable TE
material systems, such as Bi2Te3,4, 5 PbTe,6 AgSbTe2,7 Zn4Sb3,8 Zintl phase9 and SiGe,10,11 as
electrical transport properties ( For instance, the maximum power factors for n type
Hf0.5Zr0.5NiSn0.99Sb0.01 and p-type FeNb0.95Ti0.05Sb reach 50×10-4 and 106×10-4 W m-1 K-2,23,
24
respectively, which outperforms the PbTe based materials6 and Na doped SnSe.25
Moreover, HH materials possess high thermal stability and superior mechanical strength over
most other TE materials.26, 27 Therefore, HH materials have become the focus of increasing
interest in thermoelectric research.
HHs have a potential to reach higher ZTs due to the very high power factor if their intrinsic
high thermal conductivity (~10 W m-1 K-1 at 300 K)28 dominated by phonons can be
significantly reduced. Phonon scattering can be estimated by the Callaway model based on
relaxation time approximation,29 the existing phonon scattering mechanisms in the best HH
materials is comprised of Umklapp process, grain boundary, point defect and
electron-phonon scatterings. However, despite of the reduction in κL obtained by all above
mechanisms,30, 31 there is still a large room for further κL reduction given the amorphous limit
of κmin is ~ 1 W m-1 K-1 for typical HH TE materials calculated by Cahill’s formulation.32, 33
To achieve this goal, it seems that new phonon scattering mechanism is needed for HH
materials. It is important to note that the existing scattering mechanisms in HH materials
target mostly at the high frequency phonons through Umklapp and electron-phonon
scatterings, the substantial portion of phonons in the low to middle frequency (≤ 30THz)
range have been overlooked. Therefore, a new mechanism, such as dislocation arrays at the
grain boundaries,4 might be efficient in scattering low to middle frequency phonons.
Figure 1. Schematic illustration of the synthesis process of half-Heusler materials based on displacement
Seven typical HH thermoelectric materials were synthesized by the new approach, and
for each composition, three samples have been made with the sintering time of 5 minutes, 20
minutes, and 4 hours, respectively. The obtained bulk samples were firstly characterized by
powder XRD, typically for the samples being sintered for 20 minutes, as shown in Fig. 2a.
Meanwhile, in order to check the loss of materials due to liquid expelling in the SPS
process, we use the inductively coupled plasma mass spectrometry (ICP-MS) to confirm the
accurate elemental compositions. The results indicate that the actual compositions of the
(NTFC-20min) and (HZNSS-20min) bulk samples are consistent with the stoichiometric ratio
of the starting materials within the standard deviation of 1-3% (Table S1). Furthermore,
energy dispersive X-ray spectroscopy (EDX) mapping analysis indicate that the
compositional elements distributed uniformly, as shown in Fig. 2b and 2c. In Fig. 2b, Fig. 2c,
and Fig. S2, from the scanning electron microscopy (SEM) images of the cross sections, we
can see that the grains are relatively compact, while the grain size are around 150-400 nm for
NTFC-20min, and 500 nm in average for HZNSS-20min. These grain sizes and morphologies
are similar to those conventionally sintered samples reported in the literature,30 the details of
Figure 2. (a) Room temperature powder X-ray diffraction patterns for seven half-Heusler samples. The inset
shows the crystal structure model of half-Heusler and the corresponding space group (S.G.); (b) SEM images and
EDX mappings of the compositional element distribution for Nb0.8Ti0.2FeSb (NTFC-20min); and (c)
Hf0.25Zr0.75NiSn0.97Sb0.03 (HZNSS-20min).
Fig. 3a and 3b display the low-magnification TEM image of NTFC-20min and the
enlarged view of boxed region in (a), respectively. Dislocation arrays with the periodic misfit
spacing of 2.5 nm were observed in high resolution TEM image (Fig. 3c), which can be
viewed as a result of the compensation effect for the misorientation between two adjacent
grains on the (010) atomic planes along the dislocation arrays marked in red in Fig. 3d (as the
inverse fast Fourier-transformed IFFT image for 3c). In addition, the Burgers vector of each
dislocation in Fig. 3d is identified as BD = <010>. Another kind of dislocation array is shown
in the transitional Mioré patterns in HRTEM image of Fig. 3e, which has the periodic misfit
spacing of 5 nm. The IFFT and FFT images of Fig.3e are shown in Fig. 3f, and the upper
inset of Fig. 3f, respectively. The dislocation arrays in Fig. 3f are shown in red circles, and
the boxed region shows the enlarged view of the dislocation. The zone axis is identified as
[100] for both adjacent grains in Fig. 3f and the atomic planes are (010). The dislocation
arrays featured by transitional Mioré patterns are observed as the major dislocation at the
grain boundaries for NTFC-20min, which may reflect the higher crystal symmetry of HH
materials ( ̅ ) than that of the Bi0.5Sb1.5Te3 structure ( ̅ ). As expected, the dense
dislocation arrays are also observed in the HZNSS-20min sample, as shown in Fig. 3h, 3j,
and 3k, while 3h and 3j show the enlarged HRTEM images for the boxed areas of Fig. 3g and
3i, respectively. Transitional Mioré patterns are observed in Fig. 3j and the enlarged image in
To the best of our knowledge, bulk HH TE materials exhibiting dense dislocation arrays
at the grain boundaries have not been reported. 23, 30 The formation mechanism of the
dislocations in NTFC-20min, HZNSS-20min, and others samples should be similar to those
observed in materials possessing enhanced mechanical strength,37 magnetism properties,38
and TE performance.39 During the SPS process, the MgX2/LiX byproducts produced in the
high energy ball milling process can be liquefied and expelled at elevated temperatures above
873 K. The existence of liquid MgX2/LiX phase facilitates diffusion at the grain boundaries
of HHs, which upon heating and compacting, leads to the appearance of dense dislocation
arrays at grain boundaries, resulting in a semi-coherent interface structure for the HH
material.
NTFC-20min; (b) enlarged view of the boxed region in (a); (c) high resolution TEM image of a grain boundary;
(d) inverse FFT image of (c), which shows the dislocation arrays as indicated by the red symbols, inset is the
enlarged view of a dislocation; (e) high resolution TEM image of another grain boundary; (f) inverse FFT image
of (e), which shows some dislocations encircled in red, insets are the enlarged view of a dislocation (lower left)
and FFT image (upper) of the right grain; (g)-(l) TEM images featured with the dense dislocation arrays at the
The temperature dependence of the total thermal conductivity tot for six samples of
Nb0.8Ti0.2 FeSb and Hf0.25Zr0.75Ni1.05Sn0.97Sb0.03, as shown in Fig. 4a, are calculated from tot =
DCpd, where D is the thermal diffusivity coefficient (Fig. S3a), Cp the specific heat capacity
conductivity e for all samples, which was determined via the Wiedemann-Franz relation e =
LσT, where L, σ, and T is the Lorenz number (Fig. S4), temperature dependent electrical
conductivity, and the absolute temperature, respectively.40 The temperature dependent lattice
thermal conductivity L and bipolar thermal conductivity bip for the six Nb0.8Ti0.2 FeSb and
Hf0.25Zr0.75Ni1.05Sn0.97Sb0.03 samples were calculated by subtraction e from tot. The
contribution of bip arises slightly at high temperature for Nb0.8Ti0.2 FeSb. As shown in Fig. 4c
and 4d, the temperature dependence of L for NTFC-5min, NTFC-20min, HZNSS-5min, and
HZNSS-20min are significantly reduced,particularly by 40% and 44% at 300 K for the 5
minutes sintered samples in comparison to the literature values,21, 23 respectively. For the
HZNSS-5min sample, the L approaches ~1 W m-1 K-1 at 900 K, which is the minimum
lattice thermal for HH materials. Whereas, the L of 4 hours sintered samples increase
substantially and become comparable to the literatures,21, 23 which can be ascribed to the
growth of grain sizes and annihilation of dislocations, as shown in TEM images in Fig. S5.
We noticed that from 5min to 20min and 4hours, the bipolar effect in HZNSS samples
became slightly obvious, which is probably due to the slight increase of the concentrations of
the minor carrier originated from the intrinsic antisite defects between Zr and Sn sites, and in
turn the band narrowing effect in HZNSS materials.41 The reductions on L reveal that the
unique dislocation arrays at the grain boundaries might have profound impact on the phonon
scattering mechanism for these samples. In order to understand the details of the phonon
scattering mechanisms, the lattice thermal conductivity L was analyzed using the expression
based on Callaway model as following:29
∫ (1)
where , D, and is the phonon velocity, the Debye temperature, and the phonon frequency,
respectively, x is defined as x = ħ/kBT, total (x) is the total effective relaxation time, which
can be determined from the individual relaxation time originated from different scattering
However, in this work, due to the dense dislocations at the grain boundaries in the
Nb0.8Ti0.2 FeSb and Hf0.25Zr0.75Ni1.05Sn0.97Sb0.03 bulk samples, additional considerations have
to be put forth regarding the phonon scattering from the dislocation cores (DC) and the
dislocation strain (DS). The details of the calculation on the total phonon relaxation time
(τtot), as well as the individual U, EP, PD, B, DC, and DS can be found in the supporting
information.
Based on the calculations, phonon relaxation times from different scattering mechanisms
are calculated and shown in Fig. 4e and 4f for NTFC-20min and HZNSS-20min samples,
respectively. The results show that dislocation strain is the dominant mechanism for low and
middle frequency phonon scattering, and meanwhile the point defect and electron-phonon
scatterings are the dominant mechanisms for high frequency phonons in both materials. The
comparison between experimental L for six Nb0.8 Ti0.2 FeSb and Hf0.25Zr0.75NiSn0.97Sb0.03
samples and the fitted L based on the parameters from NTFC-20min and HZNSS-20min are
shown in Fig. 4g and 4h, respectively. For NTFC-5min, NTFC-20min, HZNSS-5min and
HZNSS-20min, the dislocation strain scattering is believed to contribute significantly on the
reduction of lattice thermal conductivity below 800 K, and among them the experimental and
fitting data of the 20 minutes sintering samples match well with each other. For
Nb0.8Ti0.2 FeSb samples, it is noted that the experimental thermal conductivity start to arise
slightly above 800 K, which indicates that the minority carriers could have been excited and
bipolar thermal conduction emerges. For sample NTFC-4h and HZNSS-4h, significant grain
growth accompanied by annihilation of dislocation defects at the grain boundaries can be
κe(Wm-1K-1)
HZNSS-5min
κtot(Wm-1K-1)
6 HZNSS-20min 2
HZNSS-4h
4 1
2 0
400 600 800 1000 400 600 800 1000
T(K) T(K)
(c) 5 (d) 4
NTFS-5min HZNSS-5min
NTFS-20min
κL+κbip(Wm-1K-1)
κL+κbip(Wm-1K-1)
4 HZNSS-20min
NTFS-4h HZNSS-4h
3
2
2
1 1
400 600 800 1000 400 600 800 1000
T(K) T(K)
(e) point defect Umklapp
(f) point defect Umklapp
101
Phonon relaxation time (ns)
10-1 10-1
10-3 10-3
NTFS-5min HZNSS-5min
NTFS-20min HZNSS-20min
4 NTFS-4h HZNSS-4h
3
3
2
2
1
400 600 800 1000 400 600 800 1000
T(K) T(K)
electron thermal conductivity e of all samples; (c) lattice thermal conductivity and bipolar thermal conductivity
(L+bip) of NTFC-5min, NTFC-20min, and NTFC-4h samples; (d) (L+bip) of HZNSS-5min, HZNSS-20min,
and HZNSS-4h samples; (e) calculated phonon relaxation time versus frequency for NTFC-20min sample and
for (f) HZNSS-20min sample; (g) comparison between experimental and the fitted L for three Nb0.8Ti0.2FeSb
respectively. For both compositions, the and S increase with temperature and saturate at
around 700 K for Hf0.25Zr0.75NiSn0.97Sb0.03, which is consistent with the literature results for
the similar compositions and also indicates the typical electrical transport behavior for
degenerated HH semiconductors.21, 23 It is noted here that with respect to the reported data for
experimental H ~ 29 cm2 V-1 s-1 is higher than the fitting value of 19 cm2 V-1 s-1 in the curve
under the same carrier concentration of 3.11020 cm-3, which obviously means the Sb content
is less than 0.03. This is because the difference between the experimental and fitting values
was supposed to be induced by the Sb content difference between actual and nominal values.
Homogeneous doping or alloying is particularly important for an alloying scattering
dominated carrier transport system, such as in ZrNiSn based materials since the relatively low
deformation potential and alloy scattering potential favor high carrier mobility. 42 The
relationship of μH ~ T-0.5 for HZNSS-20min shown in Fig. 5d indicates alloying scattering
mechanism for the system and this is further confirmed by the modeling in Fig. 5f. The
details of the modeling can be found in the supporting information.
Here the question still arises why the PF in HZNSS-5min and HZNSS-20min is lower than
those samples with similar carrier concentration in the literature? Despite of the scattering
induced by the minor inhomogeneity of Sb doping atoms in the sample, more contributions
may come from the crystal defects at the nanoscale level in the bulk sample, which would
scatter phonon and electron at the same time, given that the mean free paths for electron and
phonon are both revealed to be around 5 nm above 200 K for the ZrNiSn based materials. 41
The aforementioned dense dislocation strains and dislocation cores in the HZNSS-5min and
HZNSS-20min bulk samples might be the key reason. As can be seen from HZNSS-4h, with
the grain growth and annihilation of strain defects at the grain boundaries compared to two
On the contrary to ZrNiSn based materials, the p-type Nb0.8Ti0.2 FeSb system is
characterized by acoustic scattering dominated carrier transport properties, which is
consistent with our analysis in the following context. In comparison with the literature data
for the same composition,21 the Nb0.8Ti0.2FeSb bulk samples in our synthesis exhibit some
interesting features in terms of its electrical transport properties. First of all, among three
samples the temperature dependent electrical resistivity ρ (Fig. 5a) decrease with the
increasing of sintering time from 5 minutes to 4 hours, and their differences become
significant at high temperatures, eventually NTFC-4h exhibits the lowest ρ and it reaches to
9.23 μΩ m at 925 K, which is higher than the literature’s value of 7.15 μΩ m.21 Second, as for
the S, there is not as much difference as that in the ρ for three samples except in high
temperature range, where the S of NTFC-4h reaches 187 μV/K at 925 K, nearly 14 % higher
than those of NTFC-5min and NTFC-20min. Moreover, in comparison to the literatures, the S
of three samples are higher at 300 K (80 vs. 70 μV/K), and gradually coalesce with the
literature values until 925 K (184 μV/K from Ref.18). This leads to 56% increase on the PF
at 300 K for NTFC-4h sample with respect to literature, 21 reaches 39×10-4 W m-2 K-1 in our
case, as shown in Fig. 4g. Meanwhile, the PF of our sample gradually increase with
temperature until reaches 47 ×10-4 W m-2 K-1 at 535 K, than decreases slowly to 37 ×10-4 W
m-2 K-1 at 925 K, in comparison to 45 ×10-4 W m-2 K-1 at 925 K from literature. 21 The
temperature dependence of carrier concentration and Hall mobility of NTFC-20min are
shown in Fig. 5c and d, in which the carrier concentration of 201020 cm-3 is lower than the
literature by 20%, 21 probably reflecting inhomogenous Ti doping in this sample. On the other
hand, the mobility start from 17 cm2 V-1 s-1 at 300 K, then degrade in a faster way to 5 cm2 V-1
Seebeck coefficient S(T) of all samples; (c) carrier concentration nH, and (d) Hall mobility H for NTFC-20min
and HZNSS-20min; (e) and (f) are the room temperature Hall mobility as the function of carrier concentration for
NTFC-20min and HZNSS-20min, respectively. The curves were calculated by using m* = 6.9 me for NTFC-20min
and 3.05 me for HZNSS-20min; (g) power factor PF for all samples of Nb0.8Ti0.2FeSb and
Hf0.25Zr0.75NiSn0.97Sb0.03; (h) modulated PF as the function of carrier concentration for NTFC-20min and
HZNSS-20min at 300 K with experimental data points for NTFC-20min, NTFC-4h, HZNSS-20min and
The temperature dependence of ZTs for the six Nb0.8Ti0.2 FeSb and
Hf0.25Zr0.75NiSn0.97Sb0.03 bulk samples are calculated based on the measured ρ, S, and κ
values, and plotted in Fig. 6a, 6b. The TE properties of Hf0.44Zr0.44Ti0.12CoSb0.85Sn0.15 are also
put in Fig. S6. Overall, among the Nb0.8Ti0.2 FeSb samples, the ZTs increase with the sintering
time, particularly at high temperatures, where the maximum ZT of NTFC-4h reaches 0.73 at
925 K. The ZTs between room temperature and 700 K for all Nb0.8Ti0.2FeSb samples are
about 30% higher than the literature value. 21 Meanwhile, above 750 K the literature’s ZT
gradually outperforms our samples.21 These results are fully consistent with the characters of
the interplays between temperature dependence of thermal conductivity and the power factors
for the Nb0.8Ti0.2FeSb samples. Namely the enhanced PF and reduced thermal conductivity at
lower temperatures increase the ZTs over the references,21 while the lower PF at higher
temperatures overturns the reduction in thermal conductivity, leading to lower ZTs,
particularly in NTFC-5min, compared to the literature values.21 Inset in Fig. 6b shows the
carrier concentration dependence of calculated ZT based on the two-band model (conduction
band and valence band) for NTFC-20min at 300, 600, 800 K. The measured ZTs of
NTFC-20min and NTFC-4h at three different temperatures locate closely to the maximum
positions in the curves, meaning that the carrier concentration and ZTs could have been
Hf0.25Zr0.75NiSn0.97Sb0.03 samples, the calculated respective ZTs as the function of carrier concentration are shown
in the insets. The solid lines were calculated based on the two-band model. The calculated cumulative temperature
(T) dependent TE properties for all Hf0.25Zr0.75NiSn0.97Sb0.03 and Nb0.8Ti0.2FeSb samples are plotted and
compared with the literatures values for the same compositions, 21, 23 the cold side temperature is fixed as Tc = 323
K for all analysis. (c) T dependence of (ZT)eng, (d) T dependence of , (e) T dependence of (PF)eng and (f) T
Pd, and are calculated for both Nb0.8Ti0.2 FeSb and Hf0.25Zr0.75NiSn0.97Sb0.03 materials and
compared with the closest composition in the literatures.21, 23
The (ZT)eng, , (PF)eng, and Pd were calculated based on the experimental results and
compared with the data calculated from the literatures’ values.21, 23 It is shown in Fig. 6c and
6d that the temperature difference dependence of (ZT)eng and of three Nb0.8Ti0.2FeSb
samples are substantially higher than that of the reference data, with a maximum of 7.8 %
for NTFC-4h at T~600 K. Regarding the temperature difference dependence of (PF)eng and
Pd of the Nb0.8Ti0.2FeSb samples as shown in Fig. 6e and 6f, the NTFC-4h outperforms
NTFC-5min, NTFC-20min and the reference data in the whole range, owing to its much
enhanced PF at low temperatures. Meanwhile, for all Hf0.25Zr0.75NiSn0.97Sb0.03 samples, the
temperature difference dependence of (ZT)eng and as shown in Fig. 6c and 6d are
competitive to reference data with a maximum of 10.5 % at T~600 K. However, the
temperature difference dependence of (PF)eng and Pd were outperformed by reference data as
shown in Fig. 6e and 6f, which also exhibits the trend of improvement with the increase of
sintering time for the samples.
0.75 45 15
PF (10 Wm K )
9
-2
Pd (Wcm )
-1
-2
(%)
10
ZT
0.50 30 6
-4
0.25 15 5
3
FeNb0.56V0.24Ti0.2Sb FeNb0.56V0.24Ti0.2Sb
0.00 0
0 0
400 600 800 1000 0 200 400 600
T (K) T (K)
Figure
7. Thermoelectric performance for FeNb0.56V0.24Ti0.2Sb: (a) temperature dependence of ZT represented by red
filled circle, and power factor represented by blue filled squire; (b) T dependence of efficiency and power
density Pd, represented by the red filled hexagon and blue filled diamond, respectively.
In order to increase the (ZT)eng and for the p type Nb0.8Ti0.2FeSb and make them
comparable to the Hf0.25Zr0.75NiSn0.97Sb0.03 materials for better devices, the vanadium doped
samples were prepared by the similar procedures, which was enlightened by outstanding
performance of vanadium doped materials in Zhu’s report.45 The temperature dependence of
thermoelectric performance for FeNb0.56V0.24Ti0.2Sb sample is shown in Fig.7. The details of
the temperature dependence of tot, , and S for this sample can be found in Fig. S7. Fig. 7a
shows temperature dependence of ZT and PF, represented by red filled circle and blue filled
square, respectively. It is noticed that although the PF were lower than that of Nb0.8Ti0.2FeSb
owing to the decreased hole mobility and electrical conductivity, the ZT curve in convex
upward shape reaches a peak of 0.92 at 925 K, favoring overall enhanced as shown in Fig.
7b. As a result, the maximum T dependence of reaches to 10.5 % at T~600 K
accompanied by Pd of 15 W cm-2, which is just slightly lowered than the Nb0.8Ti0.2 FeSb.
3. Conclusion.
Methods
Synthesis. Six typical half-Heusler materials were synthesized through the combination of
mechano-chemical reaction and spark plasma sintering (SPS) process. All treatment and
processing of the material were carried out in an argon-filled glove box to avoid oxidation.
High purity metal chloride powders NbCl5 (99.9%), TiBr4 (98%), HfCl4 (99.9%), ZrCl4
(98%), FeCl2 (99.5%), VCl5 (99.5%), CoCl2 (97%), NiCl2 (98%), SbCl3 (99%), SnCl2 (98%),
and metal Mg/Li (99.9%) granules from Alfa Aesar were used for the synthesis. Raw
materials with a total weight of 5 g and appropriate ratios corresponding to the stoichiometric
ratio of HH materials were transferred into the stainless ball milling jar. High energy ball
milling process was then performed for 2 – 5 hrs by using a SPEX Sample Prep 8000
Mixer/Mill. The finally produced fine powder that is composed by the mixture of transitional
metal nanoparticles and MgCl2/LiCl were carefully loaded into the graphite die in the glove
box, and then treated by SPS process at 850-950 oC in the vacuumed chamber under 50 MPa
(diameter) 2 mm (thickness). The measured density of all the pellets was up to 94~98% of
the theoretical density. The obtained samples were further characterized by XRD, SEM and
TEM techniques.
Powder X-ray Diffraction (XRD). Powder XRD patterns were collected on a diffractometer
(PANalytical X’Pert Pro) in Bragg-Brentano reflection geometry with Cu K radiation
(=1.5418 Å) and operated at 40 kV and 40 mA. The patterns were carried out in the 2
range of 10 o -90 o with a step size of 0.02o at room temperature in air.
Microstructure analysis. The morphologies and element distribution of bulk samples were
characterized by a scanning electron microscopy (SEM) and energy dispersive X-ray
spectroscopy (EDX) on the S-5200 field emission SEM (Hitachi). The chemical compositions
were further checked by Inductively Coupled Plasma Mass Spectrometry (ICP-MS, X Series
2, Thermo Fisher Scientific). The microstructure and crystal defects at the atomic and lattice
levels in the bulk samples were investigated by a transmission electron microscope (TEM,
F20, FEI).
temperature dependence of thermal conductivity was calculated from the equation, = DCpd,
where D, Cp, and d are the thermal diffusivity, specific heat capacity, and density,
respectively. The thermal diffusivity was measured by the laser flash method with a
commercial system (LFA-1000, Linseis). The specific heat capacity was measured by
differential scanning calorimeter (TA Q200). The density was determined by the Archimedes
method.
1. S. Challipalli, F. H. Froes, Mechanical Alloying of Titanium-Base Alloys. Adv Mater. 1993, 5, 96-106.
2. P. V. Liddicoat, X. Z. Liao, Y. Zhao, Y. Zhu, M. Y. Murashkin et al. Nanostructural hierarchy increases the
strength of aluminium alloys. Nat Commun, 2010, 1, 63.
3. J. R. Sootsman, D. Y. Chung, M. G. Kanatzidis, New and Old Concepts in Thermoelectric Materials. Angew
Chem-Int Edit, 2009, 48, 8616-8639.
4. I. S. Kim, K. H. Lee, H. A. Mun, H. S. Kim, S. W. Hwang et al. Dense dislocation arrays embedded in grain
boundaries for high-performance bulk thermoelectrics. Science, 2015, 348, 109-114.
6. K. Biswas, J. Q. He, I. D. Blum, C. I. Wu, T. P. Hogan et al. High-performance bulk thermoelectrics with
all-scale hierarchical architectures. Nature, 2012, 489, 414-418.
7. K. F. Hsu, S. Loo, F. Guo, W. Chen, J. S. Dyck et al. Cubic AgPbmSbTe2+m: Bulk Thermoelectric Materials with
High Figure of Merit. Science, 2004, 303, 818-821.
8. T. Caillat, J.-P. F., A. Borshchevsky. Preparation and thermoelectric properties of semiconducting Zn4Sb3. J
Phys Chem Solids, 1997, 58, 1119-1125.
10. G. Joshi, H. Lee, Y. C. Lan, X. Wang, G. H. Zhu et al. Enhanced thermoelectric Figure-of-Merit in
nanostructured p-type silicon germanium bulk alloys. Nano Lett, 2008, 8, 4670-4674.
11. X. W. Wang, H. Lee, Y. C. Lan, G. H. Zhu, G. Joshi et al. Enhanced thermoelectric Figure-of-Merit in
nanostructured n-type silicon germanium bulk alloys. Appl. Phys. Lett, 2008, 93, 193121.
12. H. L. Liu, X. Shi, F. F. Xu, L. L. Zhang, W. Q. Zhang et al. Copper ion liquid-like thermoelectrics. Nat Mater,
2012, 11, 422-425.
13. W. Y. Zhao, Z. Y. Liu, P. Wei, Q. J. Zhang, W. T. Zhu et al. Magnetoelectric interaction and transport
behaviours in magnetic nanocomposite thermoelectric materials. Nature Nanotech, 2017, 12, 55-60.
15. L. D. Zhao, S. H. Lo, Y. S. Zhang, H. Sun, G. J. Tan et al. Ultralow thermal conductivity and high
thermoelectric figure of merit in SnSe crystals. Nature, 2014, 508, 373-378.
16. L. D. Zhao, J. Q. He, D. Berardan, Y. H. Lin, J. F. Li et al. BiCuSeO oxyselenides: new promising
thermoelectric materials. Energy Environ Sci, 2014, 7, 2900-2924.
17. W. Liu, X. J. Tan, K. Yin, H. J. Liu, X. F. Tang et al. Convergence of Conduction Bands as a Means of
Enhancing Thermoelectric Performance of n-Type Mg2Si1-xSnx Solid Solutions. Phys Rev Lett, 2012, 108, 5.
18. W. S. Liu, H. S. Kim, S. Chen. Q. Jie, B. Lv et al. n-Type thermoelectric material Mg2Sn0.75Ge0.25 for high
power generation. PNAS, 2015, 112, 3269-3274.
19. G. D. Mahan, Figure of merit for thermoelectrics. Journal of Applied Physics, 1989, 65, 1578.
20. S. Öğüt, K. M. Rabe, Band gap and stability in the ternary intermetallic compounds NiSnM(M=Ti,Zr,Hf): A
first-principles study. Phys Rev B, 1995, 51, 10443-10453.
21. C. Fu, T. Zhu, Y. Liu, H. Xie, X. Zhao, Band engineering of high performance p-type FeNbSb based
half-Heusler thermoelectric materials for figure of merit zT>1. Energy Environ Sci, 2015, 8, 216-220.
22. Y. Z. Pei, H. Wang, G. J. Snyder, Band Engineering of Thermoelectric Materials. Adv Mater, 2012, 24,
6125-6135.
23. S. Chen, K. C. Lukas, W. S. Liu, C. P. Opeil, G. Chen et al. Effect of Hf Concentration on Thermoelectric
Properties of Nanostructured N-Type Half-Heusler Materials HfxZr1-xNiSn0.99Sb0.01. Adv Energy Mater, 2013,
3, 1210-1214.
24. R. He, D. Kraemer, J. Mao, L. P. Zeng, Q. Jie et al. Achieving high power factor and output power density in
p-type half-Heusler Nb1-xTixFeSb. PNAS, 2016, 113, 13576-13581.
25. L. D. Zhao, G. Tan, S. Hao, J. He, Y. Pei et al. Ultrahigh power factor and thermoelectric performance in
hole-doped single-crystal SnSe. Science, 2015, 351, 141-144.
26. G. Rogl, A. Grytsiv, M. Gürth, A. Tavassoli, C. Ebner et al. Mechanical properties of half-Heusler alloys. Acta
Materialia, 2016, 107, 178-195.
27. R. He, S. Gahlawat, C. F. Guo, S. Chen, T. Dahal et al. Studies on mechanial properties of thermoelectric
materials by nanoindentation. Physica Status Solidi (a), 2015, 212, 2191-2195.
29. J. Callaway, Model for Lattice Thermal Conductivity at Low Temperatures. Phys Rev, 1959, 113,
1046-1051.
30. C. Fu, H. Wu, Y. Liu, J. He, X. Zhao et al. Enhancing the Figure of Merit of Heavy-Band Thermoelectric
Materials Through Hierarchical Phonon Scattering. Advanced Science, 2016, 1600035.
31. C. Yu, T.-J. Zhu, R.-Z. Shi, Y. Zhang, X.-B. Zhao et al. High-performance half-Heusler thermoelectric
materials Hf1−xZrxNiSn1−ySby prepared by levitation melting and spark plasma sintering. Acta Materialia,
2009, 57, 2757-2764.
32. D. G. Cahill, S. K. Watson, R. O. Pohl, Lower limit to the thermal-conductivity of disorderd crystals. Phys
Rev B, 1992, 46, 6131-6140.
34. Z. W. Chen, B. H. Ge, W. Li, S. Q. Lin, J. W. Shen et al. Vacancy-induced dislocations within grains for
high-performance PbSe thermoelelctrics. Nature Comm, 2017, 8, doi:10.1038/ncomms13828.
35. F. Casper, T. Graf, S. Chadov, B. Balke, C. Felser, Half-Heusler compounds: novel materials for energy and
spintronic applications. Semicond Sci Tech, 2012, 27, 063001.
36. F. J. Humphreys, M. Hatherly, Recrystallization and Related Annealing Phenomena. Elsevier, Oxford(2004)
37. M. Nader, F. Aldinger, M. J. Hoffmann, Influence of the α/β-SiC phase transformation on microstructural
development and mechanical properties of liquid phase sintered silicon carbide. J Mater Sci, 1999, 34,
1197-1204.
38. P. J. Jorgensen, Liquid-phase sintering of SmCo5. Journal of Applied Physics, 1973, 44, 2876.
39. H. S. Kim, S. D. Kang, Y. L. Tang, R. Hanus, G. J. Snyder, Dislocation strain as the mechanism of phonon
scattering at grain boundaries. Materials Horizons, 2016, 3, 234-240.
40. M. Jonson, G. D. Mahan, Mott Formula for The Thermopower and The Widemann-Franz Law. Phys Rev B,
1980, 21, 4223-4229.
41. T. Zhu, C. Fu, H. Xie, Y. Liu, X. Zhao, High efficiency Half-Heusler thermoelectric materials for energy
harvesting. Adv Energy Mater, 2015, 5, 1500588.
43. T. M. Tritt, Thermoelectric Phenomena, Materials, and Applications. Annu Rev Mater Res, 2011, 41,
433-448.
44. H. S. Kim, W. Liu, G. Chen, C. W. Chu, Z. F. Ren, Relationship between thermoelectric figure of merit and
energy conversion efficiency. Proceedings of the National Academy of Sciences of the United States of
America, 2015, 112, 8205-8210.
45. C. G. Fu, Y. T. Liu, X. B. Zhao, T. J. Zhu, Are solid solution better in FeNbSb-based thermoelectrics? Adv.
Electron. Mater, 2016, 1600394.
Acknowledgements
We acknowledge the funding support from the National Natural Science Foundation of China
(NSFC) under the grant number of 51572287, U1601213 and 51472052, and the financial
support from Institute of Physics at Chinese Academy of Sciences. The work performed at
the University of Houston is funded by “Solid State Solar Thermal Energy Conversion Center
(S3TEC)”, an Energy Frontier Research Center funded by the U.S. Department of Energy,
Office of Science, Office of Basic Energy Science under award number DE-SC0001299. We
would thank Professor Tiejun Zhu and Mr. Yintu Liu at Zhejiang University for their
assistance in the electrical transport property measurement for the p-type NdFeSb based
materials. We also thank Professors Lidong Chen and Xun Shi at Shanghai Institute of
Ceramics, Chinese Academy of Sciences for their insightful suggestions.
Author contributions
This paper was written by H.Z.Z. and S.M.L, and edited by G.J.S., Z.F.R., and X.L.C.
Sample synthesis, structural characterization and thermoelectric transport property
Additional information
Supplementary information is available in the online version of the paper. Reprints and
permissions information is available online. Correspondence and requests for materials
should be addressed to H. Zhao.