P. Hermet*a,
M. M. Kozab,
C. Ritterb,
C. Reibela and
R. Viennois*a
aInstitut Charles Gerhardt Montpellier, UMR-5253 CNRS, Université de Montpellier, Place E. Bataillon, 34095 Montpellier Cédex 5, France. E-mail: patrick.hermet@univ-montp2.fr; romain.viennois@univ-montp2.fr
bInstitut Laue Langevin, 6 Rue Jules Horowitz, B.P. 156, 38042 Grenoble Cédex 9, France
First published on 6th October 2015
This article is devoted to the thermal expansion of ZnSb combining experiments (neutron and X-ray) and calculations based on density functional theory. Related properties are also studied such as: the zone-center (Raman and infrared) phonon modes, the dielectric (electronic and static) tensors, the phonon density-of-states, the specific heats and the isotropic atomic displacement parameters. Our experimental data show highly anisotropic thermal expansion with large values along the a-direction. Concomitantly, a large increase of the Zn–Zn intra-ring distances and of one of the intra-ring Zn–Sb distances is observed, while other interatomic distances do not significantly change. In agreement with our calculations, the thermal expansion has positive values along the three crystal directions except around 30 K where it has weak negative values along the b and c-directions. This anomalous expansion is more important along the c-direction and it is mainly due to phonon modes with frequencies up to 75 cm−1. These modes are located in the S–Y–Γ (resp. Γ–Z) q-point range along the b (resp. c) direction. Phonon modes located in the Γ–X and in the Y–Γ–Z q-point range with frequencies up to 175 cm−1 are responsible for the positive large thermal expansion at room temperature along the a-direction. The much reduced anisotropy of the thermal conductivity is related to the lower Debye temperatures along the b and c-directions and mainly to the small transverse sound velocity between these directions.
Among these compounds, zinc antimonide (ZnSb) has been particularly studied from the forties for its good thermoelectric properties and its use in high temperature thermogeneration of electricity.1–3,8,9 The efficiency of thermoelectric generators is related to the dimensionless figure of merit: , where α is the thermopower, ρ is the electrical resistivity, κ is the thermal conductivity and T is the temperature. More than forty years ago during the first period of active search of efficient thermoelectric materials, a ZT ∼ 0.6 at 500 K was reported for p-type ZnSb.9 In the nineties, there was renewed interest for thermoelectric materials and in 1997 a large ZT ∼ 1.3 at 670 K was found for another zinc antimonide, β-Zn4Sb3, which is a metastable p-type rhombohedral semiconductor containing a large amount of interstitial and vacancy defects.10–12 Since this time, little improvement was achieved and difficulties related to its limited stability have restricted its applications.13–15 However, several groups have recently reinvestigated, both theoretically and experimentally, the AIIBV semiconducting compounds for thermoelectric applications, and more particularly ZnSb8,16–30 for its higher stability and doping versatility (it can be n-doped) than β-Zn4Sb3. Nowadays, the ZT of ZnSb has been improved using two different routes. The first route is the nanostructuration of ZnSb by ball milling for which Okamura et al.16 observed a ZT as large as 0.9 at 550 K, notably because of the decrease of thermal conductivity by a factor ∼2. However, less good thermoelectric properties were found by another group18 induced by an additional strong reduction of the power factor. Nonetheless, the stability of specimens prepared by this route requires further long-time high-temperature studies. The second route is the doping by Ag-atoms which leads to ZT ≈ 1 above 500 K for ZnSb8,22 and ZT ≈ 1.3 at 560 K for CdSb.24 The important issues that require additional studies are the long-time stability of the Ag-doped compound and the toxicity of CdSb whose the decomposition temperature is lower by about 100 K than in ZnSb.1
Thus, to optimize the thermoelectric properties of ZnSb a better understanding of its thermal and physical properties comprising the established anisotropy effects is required. As discussed by Haussermann's group,23,28 both Zn and Sb have peculiar five-fold coordination with one-like and four-unlike neighbours. The four first Sb-neighbors around Zn-atoms are forming distorted tetrahedra which share one of their edges with that of neighboring ZnSb4 tetrahedra. The two Sb atoms shared by two neighboring tetrahedra are forming the Zn2Sb2 rings with the two Zn-atoms in the center of these distorted ZnSb4 tetrahedra (see Fig. 1). Thus, the coordination of the AIIBV semiconducting compounds is ruled by an underlying tetrahedral principle, although these tetrahedra are distorted. The chemical bondings of these compounds are therefore quite unusual. Indeed, there are two types of bonding. The first one corresponds to 2 electrons–2 centers (2e–2c) covalent bondings. There are three of such bondings: one linking the Sb-atoms between them (the so-called Sb–Sb dimer), and two linking Zn-atoms with Sb-atoms belonging to different Zn2Sb2 rings. The second one is a multicenter chemical bonding formed by electrons of the four atoms belonging to the same Zn2Sb2 rings. This picture was first proposed by Velicky et al.31 fifty years ago and was then confirmed and refined very recently with two different pictures proposed for the multicenter bonding from first principles calculations.24,28
The consequences are important for understanding the anisotropy of the physical properties and more specifically the thermal properties of AIIBV compounds. Although the thermal conductivity of ZnSb and CdSb is weakly anisotropic, several studies performed several decades ago showed that in contrast their thermal expansion is strongly anisotropic.32–34 In a very recent study, Fischer et al.23 also found highly anisotropic thermal expansion in ZnAs with the b and c-lattice parameters decreasing with increasing temperature. Nevertheless, these observations still need to be confirmed both experimentally and theoretically.
In the present paper, we report an experimental study of the ZnSb thermal expansion combining neutron and X-ray diffraction. We confirm the high anisotropy of its thermal expansion which contrasts with the low anisotropy of its thermal conductivity reported in the literature. We analyze the origin of this anisotropy and the consequences on the ZnSb thermoelectric applications. We observe weak negative values of the ZnSb thermal expansion around 30 K along the b-crystal direction with higher negative values along the c-direction. The understanding of the mechanisms at the origin of these observations is established using a theoretical support based on the coupling between the density functional theory and the quasi-harmonic approximation. The analysis of the thermal expansion of ZnSb allows us to focus on other of its properties such as: the phonon density-of-states, the isotropic atomic displacement parameters, the specific heats and the lattice dynamics. For the latter, we report in particular the calculated pressure dependence of phonon modes and we give the contribution of the polar transverse optical modes to the static dielectric constants.
For the high-temperature range, we have performed XRD on powdered specimen from the single-phase batch up to 450 K using an Empyrean diffractometer with Cu Kα1 and Kα2 wavelength, a 0.013° step, and a heater made of platinum. At 500 K, the sample reacted with the platinum heater. The lattice parameter as a function of temperature was determined from Lebail refinement using the Fullprof program. For the low-temperature range, neutron powder diffraction experiments were performed on the D2B diffractometer at Institut Laue Langevin (ILL) with the 1.594 Å wavelength and a 0.05° angle step from 3 to 300 K in an Orange Cryostat. Neutron diffraction were performed on the same three batches as for the INS experiments. This could explain the small difference observed for the c lattice parameter. Rietveld refinement of the data was performed using the Fullprof program for determining not only the lattice parameter but also the atomic positions and the isotropic atomic displacement parameters as a function of temperature (see ESI†). The thermal expansion at a temperature, T, was calculated according to: , where (ai)i=1,2,3 are the lattice parameters.
INS experiments were performed on the IN6@ILL time-of-flight (ToF) spectrometer located at the ILL. An incident neutron wavelength of 4.14 Å and the time-focusing option with a focus set to 7 meV were used. Data have been collected at 300 K within a cryostat at a helium pressure of ∼10 mbars. Standard correction procedures were applied to the data comprising background and frame-overlap corrections, normalization to vanadium standard, suppression of malfunctioning detector tubes, correction of neutron wavelength dependent efficiencies of the He-detectors and multiphonon corrections. The multiphonon corrections were carried out with the software package MuPhoCorr.36 The generalized density-of-states, G(E), was calculated according to ref. 37 and 38.
Heat capacity experiments were performed between 3.5 and 40 K with the relaxation technique using a commercial microcalorimeter from Oxford Instruments. Raman scattering experiments were performed on a T64000 spectrometer from Horiba-Jobin Yvon using: a nitrogen-cooled CCD detector, backscattering geometry, triple-monochromator configuration and the 633 nm line radiation of a He–Ne laser. The resolution was 1 cm−1 and the power on the sample was 0.3 mW for a beam radius of ∼1 μm. As the size of the crystals in our polycrystalline samples was much larger than the beam size, for each investigated zone, we were probing only one crystal. Hence, some modulation on the Raman line intensities can occur due to different random crystal orientations.
Dynamical matrix, dielectric constants, Born effective charges and elastic constants are calculated within a variational approach to density functional perturbation theory43 (DFPT). Phonon dispersion curves are interpolated according to the scheme described by Gonze et al.44 In this scheme, the dipole–dipole interactions are subtracted from the dynamical matrices before Fourier transformation, so that only the short-range part is handled in real space. We considered a 3 × 3 × 3 q-points grid for the calculation of the phonon band structure and the thermal expansion, while a denser 100 × 80 × 80 grid is used for the calculation of the phonon density-of-states and heat capacity.
Lattice parameters and atomic positions were fully relaxed using a Broyden–Fletcher–Goldfarb–Shanno algorithm until the maximum stresses and residual forces were less than 2 × 10−6 GPa and 6 × 10−5 Ha per Bohr, respectively. Our relaxed lattice parameters (a = 6.290 Å, b = 7.831 Å and c = 8.237 Å) slightly overestimate the experimental ones45 (aexp = 6.202 Å, bexp = 7.742 Å and cexp = 8.100 Å) by less than 1.7% as usually observed with GGA exchange–correlation functionals.
There are three different intra-ring interatomic distances within one Zn2Sb2 ring (see Fig. 1): two Zn–Sb distances (ds1–z2 or ds3–z4 and ds1–z4 or ds3–z2) and one Zn–Zn distance (dZ2–Z4). Their thermal variations are shown in Fig. 3. The longest S1–Z2 distance and the Z2–Z4 distance increase by about ∼0.86% upon T changes, whereas the shortest S1–Z4 distance of about 2.75 Å does not change.
![]() | ||
Fig. 3 Thermal variation of the intra-ring (top) and inter-ring interatomic distances (bottom). See Fig. 1 for the atom labels. |
Upon heating the intra-ring dS1–S3 increases by ∼0.32% from 4.91 Å at 3 K to 4.925 Å at 305 K, which is significantly less than for the case of Zn–Zn distance. Because all four atoms forming a Zn2Sb2 ring are in the same plane, the deformation of the rings can also be characterized by the interatomic angles. They are reported in details in the ESI.† It is worth to note that all the Zn–Zn–Sb angles, and thus the Z2–Z4 distance, change significantly with temperature. In the same manner, the Sb–Zn–Sb and Zn–Sb–Zn angles comprising the longest ds1–z2 or ds3–z4 vary strongly with the temperature. All the above results indicate that the shape of the Zn2Sb2 rings is deformed due to the increased Zn–Zn distance.
Let us focus now on the inter-ring distances. As can be seen in Fig. 3, there are three inter-ring distances to consider: the two shortest Zn–Sb distances, ds1–z1 or ds5–z4 and ds1–z3 or ds4–z4, and the long inter-ring Sb–Sb distance (dS1–S2). Both ds1–z1 and dS1–S2 increase by ∼0.1–0.2%, which is comparable to the experimental error. The shortest ds1–z3 increases by 0.325%. Note that some of the interatomic angles implying at least one inter-ring bonding change significantly with increasing temperature whereas some others do not. These are essentially angles comprising the shortest inter-ring distance ds1–z3. See ESI† for detailed results.
In conclusion, these results indicate that an uniform increase of the size of the Zn2Sb2 rings upon heating is related to the dominant increase of the Zn–Zn intra-ring distances. The distances between the Zn2Sb2 rings show a smaller variation while experiencing a slight rotation. The large anisotropy of the thermal expansion, and especially the large expansion along the a-direction, is linked to the strong deformation of the Zn2Sb2 rings. These trends are also observed in ZnAs as it shows23 a dominant thermal expansion along a, and similar thermal variation of the Zn–As intra-ring distances and Zn-distances. Finally, we cannot establish a clear correlation between the changes in the ZnSb microscopic structure and its weak negative thermal expansion at low temperatures along the c-direction because of the limited experimental data below 100 K.
The fixed-strain relaxed-ion dielectric tensor (εη) can be obtained by adding to ε∞ the contribution coming from the response of the ions (εph) to the electric field. To estimate this phonon-mediated contribution, one can use a model that assimilates the solid to a system of undamped harmonic oscillators. Doing so, εη appears as:43,48
![]() | (1) |
B1u-modes | B2u-modes | B3u-modes | |||||||
---|---|---|---|---|---|---|---|---|---|
ωm | Sm | εph33,m | ωm | Sm | εph22,m | ωm | Sm | εph11,m | |
TO1 | 51 | 1.978 | 1.710 | 38 | 1.591 | 2.402 | 57 | 2.615 | 1.778 |
TO2 | 54 | 0.570 | 0.426 | 56 | 0.283 | 0.196 | 62 | 1.071 | 0.613 |
TO3 | 117 | 10.168 | 1.629 | 110 | 3.183 | 0.582 | 120 | 9.959 | 1.531 |
TO4 | 142 | 20.130 | 2.208 | 140 | 0.147 | 0.017 | 154 | 23.114 | 2.156 |
TO5 | 176 | 28.711 | 2.046 | 184 | 43.523 | 2.834 | 173 | 16.745 | 1.231 |
Total (phonons) | 8.019 | 6.031 | 7.310 |
![]() | ||
Fig. 4 (Left hand side) Calculated infrared reflectivity spectra on polycrystal and on single crystals with light polarization along (100) [B3u], (010) [B2u] and (001) [B1u] directions. (Right hand side) Experimental unpolarized Raman spectra. Open symbols represent experimental data of Smirnov et al.49 Calculated positions of the Raman lines with their symmetry are sketched by the bars and crosses at the bottom of the figure. |
Mode symmetry | TO-modes | LO-modes | |||||
---|---|---|---|---|---|---|---|
Calc. (cm−1) | Exp.49 (cm−1) | dω/dP (cm−1 GPa−1) | γm | Calc. (cm−1) | |||
B2u | TO1 | 38 | 44 | −0.14 | −0.18 | LO1 | 39 |
B1u | TO1 | 51 | 58 | −0.20 | −0.19 | LO2 | 52 |
B1u | TO2 | 54 | 0.00 | 0.00 | LO2 | 55 | |
B2u | TO2 | 56 | 61 | 0.12 | 0.11 | LO2 | 57 |
B3u | TO1 | 57 | 66 | −0.13 | −0.11 | TO1 | 58 |
B3u | TO2 | 62 | 0.49 | 0.38 | LO2 | 63 | |
B2u | TO3 | 110 | 119 | 1.01 | 0.44 | LO3 | 111 |
B1u | TO3 | 117 | 121 | 1.29 | 0.53 | LO3 | 120 |
B3u | TO3 | 120 | 123 | 1.10 | 0.44 | LO3 | 122 |
B2u | TO4 | 140 | 1.55 | 0.53 | LO4 | 140 | |
B1u | TO4 | 142 | 154 | 1.78 | 0.60 | LO4 | 147 |
B3u | TO4 | 154 | 166 | 1.51 | 0.47 | LO4 | 158 |
B3u | TO5 | 173 | 184 | 1.79 | 0.50 | LO5 | 178 |
B1u | TO5 | 176 | 189 | 1.83 | 0.50 | LO5 | 184 |
B2u | TO5 | 184 | 195 | 1.94 | 0.51 | LO5 | 194 |
The experimental unpolarized Raman spectrum of ZnSb is displayed in the right hand side of Fig. 4, together with the position of the Raman lines measured by Smirnov et al.49 and obtained by our calculation. We observe a quite good agreement between our calculation and the two sets of experimental data. This agreement is even better with our experimental data as Smirnov et al.49 do not clearly observe the Raman lines between 120 and 160 cm−1. We unambiguously assign two additional experimental lines at 130 and 146 cm−1.
The pressure dependence of the zone-center phonon modes has been calculated using the mode Grüneisen parameter:
![]() | (2) |
Raman-modes | Silent-modes | ||||||
---|---|---|---|---|---|---|---|
Mode symmetry | Freq. (cm−1) | dω/dP (cm−1 GPa−1) | γm | Mode symmetry | Calc. (cm−1) | dω/dP (cm−1 GPa−1) | γm |
B2g | 48 | −0.05 | −0.05 | Au | 35 | −0.24 | −0.32 |
Ag | 54 | 0.03 | 0.02 | Au | 44 | 0.15 | 0.16 |
B1g | 61 | 0.17 | 0.13 | Au | 58 | 0.17 | 0.14 |
B3g | 75 | 0.33 | 0.21 | Au | 91 | 0.39 | 0.21 |
Ag | 75 | 0.80 | 0.51 | Au | 153 | 2.05 | 0.64 |
Ag | 81 | 0.55 | 0.33 | Au | 185 | 1.99 | 0.52 |
B3g | 85 | 0.83 | 0.47 | ||||
B1g | 87 | 1.13 | 0.62 | ||||
B2g | 99 | 1.13 | 0.54 | ||||
B3g | 114 | 1.18 | 0.50 | ||||
B1g | 119 | 0.87 | 0.35 | ||||
B2g | 121 | 1.03 | 0.41 | ||||
B2g | 144 | 1.26 | 0.42 | ||||
B3g | 155 | 1.22 | 0.38 | ||||
B1g | 158 | 0.63 | 0.19 | ||||
B3g | 160 | 0.72 | 0.21 | ||||
Ag | 161 | 0.93 | 0.28 | ||||
Ag | 165 | 1.37 | 0.40 | ||||
B2g | 167 | 0.58 | 0.17 | ||||
B1g | 172 | 1.45 | 0.40 | ||||
B1g | 177 | 1.49 | 0.40 | ||||
B3g | 183 | 1.58 | 0.41 | ||||
B2g | 189 | 1.73 | 0.44 | ||||
Ag | 190 | 1.78 | 0.45 |
Fig. 6 reports the experimental generalized density-of-states, G(E), with the calculated atom-projected and the total phonon density-of-states, where qj is the phonon wavevector q in branch j and N is the number of atom in the primitive unit cell. One can see that the main features in Z(E) are observed in G(E) obtained from INS at room temperature. We observe a small gap between the optical branches around 130 cm−1. When comparing with the experiment, this gap is associated to the dip at ∼140 cm−1 due to a decreasing resolution of the spectrometer at high energy (this progressively broadens the experimental sharp features at high energy).
Acoustic branches (LA + TA) mainly contribute to the lowest experimental peak at 44 cm−1. The peak at 84 cm−1, located at the X-high symmetry point, mainly involves the Zn-atoms moving along the a-direction. Our measurements also support the presence of a peak around 48 cm−1 (6 meV) as seen by Belash et al.50 using INS on a sample containing a mixture of ZnSb and antimony, and by Claudio et al.51 on nanometric Zn1+xSb using antimony nuclear inelastic scattering. In addition, comparing our results on bulk ZnSb with those on Zn4Sb3,30,52,53 the low energy peak in G(E) is located around the same energy (∼5.5 meV) in both compounds.
![]() | (3) |
![]() | ||
Fig. 7 Calculated specific heat at constant-volume (Cv) and constant-pressure (Cp) with the experiment (lower pannel). Mamedova et al. and Danilenko et al. refer to ref. 55 and 56, respectively. Experimental isotropic atomic displacement parameters (upper pannel). Inset: Debye presentation, Cp/T3, for experimental and calculated data. |
We get a calculated Debye temperature, θD = 217 K, from the linear fit of Cv with respect to T3 at very low temperatures (T < 4 K), in good agreement with that calculated from the elastic constants (θD = 233 K). At low temperatures, we have a peak calculated at 13 K which agrees well with our experimental heat capacity data where it is found at 15 K (see inset of Fig. 7). Such a deviation of Cp/T3 from a constant value expected for a basic Debye behaviour is related to the large peak seen in G(E)/E2 at about 47 cm−1 (see inset of Fig. 6). In both cases, this excess signal has the same origin mainly related to the presence of zone boundary acoustic modes and also of a low-energy optical mode ∼47 cm−1.
Fits of the Einstein and Debye models to the experimental data result in very similar values for the characteristic temperatures ΘE and ΘD of Zn and Sb. This highlights as well that the dynamics of Zn and Sb are strongly coupled. For the Einstein model we derive a ΘE of 128 K for Zn and 130 K for Sb. These values are about twice larger than the energy of the first peak in the G(E) at about 70 K. For the Debye model we derive a ΘD of 222 K for Zn and 224 K for Sb. They agree well with Debye temperatures derived from experimentally determined elastic constants (253 K).4
Note that we were able to model the Uiso only with the high temperature linear extrapolation of the Einstein and Debye models from 50 K to 250 K. Indeed, experimental Uiso at ∼3 K was too small for being fitted with the Einstein and Debye models. Another striking feature of Uiso is the behaviour at 300 K. The Uiso tend to higher values than the linear T-dependence predicted by a harmonic model. This is in line with the departure of the experimentally determined Cp towards higher values than predicted within harmonic approximation (see Fig. 7). For this reason only data up to 250 K were considered for the Einstein and Debye fits. Mozharivskyj et al.54 obtained from powder X-ray diffraction at room temperature even higher values of Uiso(Zn) = 0.019 Å2 and Uiso(Sb) = 0.0162 Å2.
![]() | (4) |
![]() | (5) |
Eqn (5) has been calculated from finite differences. For instance, to compute γ1, the phonon dispersion curves have been calculated at two perturbated structures that are derived from the equilibrium structure by straining the a-lattice parameter by ±1%, and reoptimizing the atomic positions. The left hand side of Fig. 8 shows the calculated dispersion of the γ(j, q) along the three crystal axes. The dispersions are discontinuous at the zone-center as a consequence of the anisotropy and polarization dependence of the sound velocities. Positive and negative γ(j, q) are observed along the three crystal directions and the whole Brillouin zone. Some mode Grüneisen parameters are weakly negative (∼−0.5) along the a-direction and are pronouncely negative along the two other directions (up to ∼−2.5). Modes with negative γ(j, q) are located up to 65 cm−1 along the a and b-directions, and up to 100 cm−1 along the c-direction (see Fig. 8, right). Above these threshold frequencies, all modes have positive Grüneisen parameters with maximal values larger than 3 near the X and Γ high-symmetry points along the a-direction while no clear maximum is observed for the two other directions. As a consequence, an anomalous negative thermal expansion could occur at low temperatures along the b and c-directions. In contrast, we expect this effect is strongly diminished at low temperatures along the a-direction or even suppressed, leading to positive αa(T).
![]() | ||
Fig. 8 Calculated mode Grüneisen parameters [γj(q)] along high-symmetry crystal directions (left hand side) and as a function of the frequency (right hand side). |
We also derived the thermodynamic Grüneisen parameters:57
![]() | (6) |
The thermal expansion measured up to 450 K is reported in Fig. 9 combining the elastic neutron scattering (2–300 K) and X-ray (300–450 K) experiments. We observe a strong anisotropy in agreement with prior works on AIIBV compounds23,32–34 and small negative values around 30 K along the b and c-directions. The negative thermal expansion is however more pronounced in the c-direction. Between 100 and 300 K, αa and αb values are close to those reported by Anatychuk and Mikhalchenko,34 while αc is twice larger. Among other AIIBV compounds, CdSb has similar αa and αc, but αb is significantly smaller.32,33 In contrast, the thermal expansion of ZnAs is bisected along a and negative along the two other directions.23 At room temperature, the volume thermal expansion of ZnSb (34 MK−1) is significantly smaller than in Zn4Sb3 where it is between 46.5 MK−1 (ref. 15) and 58 MK−1 (ref. 10 and 58).
The calculated temperature dependence of the linear thermal expansion of ZnSb is in quite good agreement with the experimental data although αc tends to higher values than the data derived from diffraction experiments above 300 K. Along the a-direction, the thermal expansion has a standard behaviour: a T3 dependence at low temperatures, followed by a monotonic increase in the intermediate temperature range, and a constant value at high temperatures. In contrast, for the two other directions, the thermal expansion does not follow this standard dependence. Indeed, the acoustic modes and some low-frequency optical modes are excited up to 30 K. These modes have in majority negative mode Grüneisen parameters, and the sign of the thermal expansion becomes negative (see Fig. 8). Above 30 K, the optical modes are beginning to be increasingly excited and the contribution of these modes with a positive Grüneisen parameter increases. This has the effect of reversing the sign of the thermal expansion that is now positive and increases monotonically to become constant at high temperatures as all optical modes are now excited.
The calculated negative thermal expansion around 30 K is mainly due to phonon modes with frequencies up to ∼75 cm−1. Along the b-direction, they are located in the S–Y–Γ q-point range and the most negative contribution comes from the silent Au mode centered at 35 cm−1. Along the c-direction, they are located between Γ and Z high-symmetry point and most negative contribution comes from the polar B2u mode centered at 38 cm−1. The eigendisplacement vectors of these two modes displayed in Fig. 10 show that the Zn and Sb intra-ring atoms have a wagging-type vibration (i.e. each atom type moves in phase and out of the ring plane) and the inter-ring atoms have a complex torsional mode. This kind of motion leads to a network contraction and therefore to a negative thermal expansion as the ZnSb structure is built from Zn2Sb2 rhomboid rings which are arranged in layers and linked to 10 neighboring rings (see Fig. 1). We think that this explanation can be applied for the understanding of the negative thermal expansion of ZnAs between 100 and 300 K as Fischer et al.23 observed a decrease of the inter-ring Zn–As bondings in this temperature range.
The calculated positive thermal expansion at room temperature along the a-direction is mainly due to phonon modes located in the Γ–X and in the Y–Γ–Z q-point range with frequencies up to 175 cm−1. In particular, it is dominated by the Raman mode at 75 cm ± 1 (Ag) and the mode at 83 cm−1 located at the X-point. These modes are assigned to a scissoring-type vibration of the Zn–Zn intra-rings along the a-direction (see Fig. 10 in the case of the Ag mode). This assignment is consistent with our analysis of the thermal dependence of the ZnSb crystallographic parameters and notably with the large increase of the Zn–Zn distances inside the Zn2Sb2 rings (see Section IV.A). So, the large anharmonicity of the Zn-motions along the a-direction is one of the key ingredients for explaining the large anisotropy of the ZnSb thermal expansion and especially its large value.
![]() | (7) |
Within the more rigorous present approach, we get an experimental thermodynamic Grüneisen parameter Γ ∼ 1.1 at 300 K (see eqn (S-3) in ESI†). This value is slightly smaller than the average Grüneisen parameter ∼ 1.2 obtained from our DFPT calculations and much larger than that reported by Bjerg et al.30 from DFT calculations
. This experimental thermodynamic Grüneisen parameter is however smaller than that obtained experimentally for Zn4Sb3 (1.35 < Γ < 1.57).10,53 This indicates that the overall anharmonicity is larger in Zn4Sb3 than in ZnSb.
Using our experimental Γ and the Debye temperature obtained from Balazyuk's work4 (θD = 253 K), the Slack model gives a thermal conductivity between 5.87 and 6.37 W m−1 K−1 at 300 K. This is more than 50% larger than the experimental values for bulk samples (2–4 W m−1 K−1).3,11,18,23,34 The agreement is however reasonable as the Slack's model usually overestimates the thermal conductivity compared to experiments.30,59 When using the DFPT calculated parameters, one finds κSlack = 3.95 W m−1 K−1 at 300 K which is close to the experimental values. Note also that Bjerg et al.30 found a larger thermal conductivity mainly due to a smaller Γ.
The thermal conductivity is weakly anisotropic as the κa:
κb
:
κc ratio is 1
:
1.05
:
1.15.3 This contrasts with the very anisotropic thermal expansion and the rather anisotropic Grüneisen parameter. Moreover, the thermal conductivity is the smallest for the direction with the largest thermal expansion and Grüneisen parameter and vice versa. In order to understand the anisotropy of the thermal conductivity, Anatychuk and Mikhalchenko34 have used a modified version of the Slack's equation in which they respectively replaced the Debye temperature θD and the thermodynamic Grüneisen parameter Γ by the Debye temperatures θD,i and the Grüneisen parameters Γi that can be indeed defined for each i-direction in the case of anisotropic system from the sound velocity vsi (see ESI† for details). However, if this is a too rough approximation to reach an understanding of the anisotropy of the thermal conductivity, we point out however that the anisotropy of the sound velocity and hence Debye temperatures θD,i, are the inverse to that of the Grüneisen parameters Γi. Indeed, the Debye temperatures for the b and c-directions are smaller by about 15% both experimentally and in DFT. This is because the transverse sound velocity vtbc between b and c-directions is much smaller than for the other directions. It seems therefore that the small anisotropy of the thermal conductivity of ZnSb could be due to a partial compensation of the anisotropy of the Debye temperatures θD,i and of the sound velocities by the opposite anisotropy of the Grüneisen parameters and hence of the anharmonicity and Umklapp relaxation time. The above discussion is just qualitative and a full explanation of the anisotropy of the thermal conductivity is beyond the scope of the present paper and would require to calculate the thermal conductivity.
For high temperature thermoelectric applications, the thermoelectric materials experience a large thermal gradient that can induce large stresses along the thermoelectric leg. The knowledge of both thermal expansion and elastic constants is useful for evaluating if the material will crack or not under large thermal gradient and under long duration. Therefore, materials with low thermal expansion are suitable. As we have seen, ZnSb has strongly anisotropic thermal expansion and it will be only weakly deformed by thermal gradient along c axis. Its average linear thermal expansion α is about 12 MK−1 at room temperature and only slightly increasing at higher temperatures. This value is comparable to that of other thermoelectric antimonide compounds such as skutterudite or clathrate compounds (α = 9–14 MK−1)60–63 but significantly larger than in silicon and germanium which have small α (∼2–6 MK−1).57,62 On the contrary, the thermal expansion is significantly lower than in Bi2Te3 used around room temperature with rather small thermal gradient and in PbTe used at larger temperature and with larger thermal gradient which have rather large thermal expansion (17 and 20 MK−1 respectively).62 As discussed by Case,62 it is necessary to have the smallest possible αE value (E being the Young modulus) and also a small Poisson coefficient in order to minimize the maximum thermal stress due to the thermal gradient and to maximize the thermal shock resistance parameter. In the case of ZnSb (with α from present work and E from Balazyuk's work4), one finds the isotropic αE = 962.7 GPa MK−1 which is typical of thermoelectric materials.62 Indeed, when comparing αE with those of other thermoelectric materials, one sees that this is 30% larger than in Bi2Te3 but 20% smaller than in PbTe and much smaller than in antimony-based skutterudites, but this is however much larger than for silicon or germanium.62
Another important point for thermoelectric applications is the interfacial stress between the thermoelectric material and their electrically contacting materials. As copper is often used as electrical contact, one could minimize the interfacial stresses if the thermal expansion of thermoelectric material was close to that of copper (17 MK−1 (ref. 57)). From this point of view, the average thermal expansion of polycrystalline ZnSb is close to that of copper but the thermal expansion of single-crystal oriented along c direction is three times smaller. Therefore, from thermal stresses point of view, probably polycrystalline samples are better suited than single-crystalline samples for high temperature applications.
To summarize, ZnSb seems to have thermo-mechanical properties at least as good as the usual thermoelectric materials.
First, our experimental data show highly anisotropic thermal expansion with large values along the a-direction. Concomitantly, a large increase of the intra-ring Zn–Zn and of one of the Zn–Sb distances is observed, while other interatomic distances do not significantly change. The thermal expansion of ZnSb is positive except around 30 K where it is weakly negative along the b-direction with a more pronounced effect along the c-direction.
Then, the calculated mode-by-mode decomposition of the thermal expansion shows that the negative expansion is mainly due to phonon modes with frequencies up to 75 cm−1. Along the b-direction, they are located in the S–Y–Γ q-point range and the most negative contribution comes from the silent Au mode centered at 35 cm−1. Along the c-direction, they are located between Γ and Z high-symmetry point and the most negative contribution comes from the polar B2u mode centered at 38 cm−1. The eigendisplacement vectors of these two modes show that the Zn and Sb intra-ring atoms have a wagging-type vibration and the inter-ring atoms have a complex torsional mode, leading to a network contraction. The calculated positive thermal expansion at room temperature along the a-direction is mainly due to phonon modes located in the Γ–X and in the Y–Γ–Z q-point range with frequencies up to 175 cm−1. In particular, the Raman mode at 75 (Ag) and the mode at 83 cm−1 located at the X-point dominate this standard thermal expansion. These modes are assigned to a scissoring of the Zn–Zn intra-rings along the a-crystal direction. This assignment is consistent with our analysis of the thermal dependence of the ZnSb crystallographic parameters and notably with the large increase of the Zn–Zn distances inside the Zn2Sb2 rings. This work provides benchmark theoretical and experimental results to understand the thermal expansion mechanisms of AIIBV compounds belonging to orthorhombic structures.
Finally, we have discussed the origin of the much reduced anisotropy of the thermal conductivity compared to the anisotropy of the thermal expansion and Grüneisen parameters. It is related to a compensation by the lower Debye temperatures along the b and c-directions and mainly to the small transverse sound velocity between these directions. ZnSb is promising for future thermoelectric applications as it has thermo-mechanical properties similar to the usual thermoelectric materials.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra16956a |
This journal is © The Royal Society of Chemistry 2015 |