Ajay Kumar
Verma
abcd,
Shamma
Jain
ab,
Kishor Kumar
Johari
ab,
Christophe
Candolfi
e,
Bertrand
Lenoir
*e,
Sumeet
Walia
cd,
S. R.
Dhakate
ab and
Bhasker
Gahtori
*ab
aCSIR-National Physical Laboratory, Dr. K. S. Krishnan Marg, New Delhi 110012, India. E-mail: bhasker@nplindia.org; Fax: +91-11-45609310; Tel: +91-11-45608556
bAcademy of Scientific & Innovative Research (AcSIR), Ghaziabad, 201002, India
cFunctional Materials and Microsystems Research Group and the Micro Nano Research Facility, RMIT University, Melbourne, VIC 3001, Australia
dSchool of Engineering, RMIT University, GPO Box 2476, Melbourne, Victoria 3001, Australia
eInstitut Jean Lamour UMR 7198 CNRS – Université de Lorraine Campus ARTEM, 2 allée André Guinier, BP 50840, Nancy 54011, France. E-mail: bertrand.lenoir@univ-lorraine.fr
First published on 27th November 2023
Half-Heusler (HH) alloys based on TiCoSb are becoming popular semiconducting materials for mid-temperature thermoelectric (TE) applications, due to their superior Seebeck coefficient, moderate electrical conductivity, and excellent mechanical properties. However, their practical applicability is mainly limited by their high lattice thermal conductivity. Here, we show how a multi-alloying approach that involves co-substitution of Zr and Hf on the Ti site and Bi on the Sb site can significantly lower the heat transport of the TiCo0.85Fe0.15Sb alloy due to enhanced mass and strain-field fluctuations in the lattice. The substantial rise in point-defect phonon scattering leads to a sharp reduction in the lattice thermal conductivity from 8 to 2 W m−1 K−1 at ∼300 K and from 5 to 1.74 W m−1 K−1 at ∼843 K in Ti0.5Zr0.2Hf0.3Co0.85Fe0.15Sb0.96Bi0.4. The achieved thermal conductivity is the lowest value reported so far among TiCoSb-based alloys. Importantly, the reduction in thermal conductivity outweighs the concomitant decrease in the power factor, ultimately leading to an improved thermoelectric figure of merit ZT. Our findings show that creating large point defects through heavy isovalent substitution is an effective approach to significantly reduce the phonon transport in HH alloys.
In addition to chalcogenides,19,20 skutterudites,21,22 silicides23,24 and clathrates25,26 extensively studied over the last decade for power generation applications, half-Heuslers (HH) alloys are promising candidates owing to their decent thermoelectric performance, low cost, composed via environmentally-friendly elements, good thermal stability and high mechanical robustness.27–30 The cubic crystal structure of HH compounds with the XYZ stoichiometry is built up by four interpenetrated face-centered cubic (fcc) lattices with one vacant position. Semiconducting properties generally emerge when the chemical composition is valence-balanced. Both n-(XNiSn; X = Ti, Zr or Hf) and p-type (XCoSb and X′FeSb; X = Ti, Zr or Hf and X′ = Nb or Ta) compositions can be designed to achieve high thermoelectric performance via proper optimization. More complex compositions such as vacancy-containing , double, triple and high-entropy HHs have been recently studied, providing interesting avenues to engineer novel compositions with improved thermoelectric performance.31–33
Among these compositions, TiCoSb-based alloys have received significant attention because of the tuning ability to develop both n- and p-type compatible counterparts with similar thermal expansion coefficients which are prerequisite for TE devices. TiCoSb has complex electronic band structure that can be engineered to increase valley degeneracy, which contributes to the high power factors achieved in p-type alloys.7 However, high κL is one of the key factors that restricts the TE performance of TiCoSb based compounds. Significant efforts were devoted to lowering κL using various approaches such as doping,34 nanostructuring35 or nano-composite.36 The possibility to parallel enhance point-defect scattering through atomic substitutions and induce electronic band convergence in p-type TiCoSb alloys have been demonstrated in recent studies, yielding decreased κL values of 2.5 W m−1 K−1 at 823 K in Ti0.8Zr0.2Co0.85Fe0.15Sb0.96Sn0.04.37 The resulting improved maximum ZT value of 0.54 at 823 K places this alloy among the best reported p-type TiCoSb-based compositions without phase separation.
In this work, we explore alternative routes to further reduce κL by creating additional point defects in p-type TiCo0.85Fe0.15Sb via heavy substitution on the Ti and Sb sites with Zr/Hf and Bi, respectively. These substitutions enhance mass and strain field fluctuations in the lattice which drive a significant decrease in κL of ∼65%, eventually reaching ∼1.74 W m−1 K−1 at 843 K for Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04. This reduced heat transport improves the ZT values by ∼60% compared to TiCo0.85Fe0.15Sb, leading to a peak ZT value of ∼0.3 at 843 K. Not only do these findings further demonstrate the benefit from heavy alloying to reduce the phonon heat transport in HH alloys but they also indicate the importance of carefully choosing the substituting elements to maintain the convergence of the valence band edges in TiCoSb alloys.
Patterns of powder X-ray diffraction (PXRD) were acquired at room temperature for each of the samples by using an X-ray diffractometer (Rigaku Miniflex II). The PXRD data have been performed for Rietveld refinements using the FullProf software.39 The field-emission scanning electron microscope (FESEM, TESCAN MAGNA GMH) and the energy-dispersive X-ray spectroscopy (EDS, EDAX OCTANE ELECT SUPER) were used to investigate the surface morphology and chemical homogeneity, respectively. In order to further investigate microstructural features at the atomic scale, high-resolution transmission electron microscopy (HRTEM; JEM 2100F-JOEL 200 kV) was performed.
Electrical resistivity and Seebeck coefficient were simultaneously measured with a ZEM-3 system (ULVAC-RIKO) in the temperature range of 300 to 823 K. The Hall carrier density pH and mobility μH were measured at room temperature using a Hall effect instrument (HEMS, Nanomagnetics). Using sapphire as a standard and a differential scanning calorimeter system (DSC-NETZSCH, 404 F3), the isobaric specific heat (cp) of each sample was determined between 300 and 773 K. The thermal diffusivity (D) of disk-shaped samples was measured by employing a laser flash method (Netzsch LFA 427). The total thermal conductivity κ (= κL + κe) was calculated using the relation κ = cpDd.
Fig. 1 (a) Room-temperature PXRD patterns and (b) Rietveld refinement performed on the illustrative sample Ti0.8Zr0.2Hf0.3Co0.85Fe0.15Sb0.96Bi0.04. |
The surface morphology and elemental chemical configuration of the illustrative sample Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04 were investigated by FESEM and EDXS as depicted in Fig. 2. The FESEM images show that the sample is densely packed with the presence of few minor secondary phases in the HH matrix. The chemical compositions (Fig. 2(b)), corresponding to the broad area shown in Fig. 2(a), is very close to the nominal stoichiometry. The minor secondary phase observed in Fig. 2(c) corresponds to a Fe-rich phase (Fig. 2(d), dark contrast region). Fig. 2(d) and (e) show the FESEM images of alumina-polished samples at low and high resolution, which confirm their high density. The elemental mapping of the representative sample x = 0.04 (Fig. S1 in ESI†) confirms the uniform distribution of the elements with only the Fe-rich phase as a minor secondary phase.
To further investigate the microstructure at the atomic scale, HRTEM analysis of the representative sample Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04 was carried out (Fig. 3). The extracted interplanar spacing of the lattice planes are in good agreement with that of the TiCoSb HH phase. Two regions, R1, and R2 were analyzed at the atomic scale. The region R1 consists of a plane of calculated d-spacings 0.2117 nm nearly corresponding to the (220) family of planes, which is more dominating in TiCoSb HH phase as found in XRD data. Moreover, the edge dislocation defects are observed in region R1, while the region R2 consists of four overlapping planes of d-spacing 0.2845 nm, 0.2117 nm, 0.1885 nm, and 0.158 nm, which are nearly corresponding to (200), (220), (311), and (400) family of planes of TiCoSb HH phase, and also show the edge dislocations. The analyses demonstrate the presence of a large distribution of edge dislocation defects, which are expected to impact the transport properties. Fig. 3(b) shows the selected area electron diffraction (SAED) pattern of the same sample, where the superimposition of bright spots with Debye rings, are clearly visible, confirming the polycrystalline nature of the material.40 The diffraction pattern reflects the material's crystal structure in the reciprocal space where Debye rings are found corresponding to the family of planes (111), (200), (220), (222), (400), (311), (422) of TiCoSb HH phase.
Fig. 4 illustrates the temperature dependence of the electronic transport parameters of all the samples Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb1−xBix (0 ≤ x ≤ 0.04). In order to facilitate a valuable comparison, the transport parameters of TiCo0.85Fe0.15Sb, which were determined in a previous experiment, have been included.7 At 300 K, σ tends to decrease in samples alloyed with Zr and Hf with respect to TiCo0.85Fe0.15Sb. Further alloying with Bi (0 ≤ x ≤ 0.04) weakly affects σ due to the small concentration of Bi introduced on the Sb site. The increasing trend of σ with increasing temperature indicates semiconducting behavior of all samples. Hall effect measurements carried out at 300 K (Table 1) reveal that the hole concentration pH drops from 5.4 to 1.2 × 1020 cm−3 as the Bi content increases. A similar behavior of decreased carrier density upon isoelectronic Bi doping on Sb has also been observed in other p-type compounds.41,42 The corresponding hole mobilities μH remain very low, varying in the range 0.68–1.57 cm2 V−1 s−1 at 300 K. These values are similar to those achieved in other heavily-alloyed HH compounds43 and can be mainly ascribed to alloy scattering that strongly limits hole mobility.
Parameters | X = 0 | X = 0.01 | X = 0.02 | X = 0.03 | X = 0.04 |
---|---|---|---|---|---|
Hole density pH (×1020 cm−3) | 5.40 | 4.68 | 1.54 | 2.59 | 1.21 |
Mobility μH (cm2 V−1 s−1) | 0.68 | 0.67 | 1.36 | 1.38 | 1.57 |
The temperature dependence of the weighted mobility (μw) was calculated using the measured σ and α:44
Fig. 4(c) displays the temperature-dependent Seebeck coefficient α for each sample. The positive value of α over the entire temperature range indicates an electrical conduction dominated by holes. The room-temperature α values significantly decrease upon Zr and Hf alloying from 137 to 80 μV K−1 in spite of the concomitant decrease in the Hall carrier density. This counterintuitive trend might be rationalized through modifications of the electronic band structure and, in particular, to an increase in the energy offset between the valence band maxima located at the Γ and L points of the Brillouin zone. The energy offset of 51 meV in TiCoSb can be manipulated by alloying on the Co and Sb sites, which modifies the energy of the band edges at the L and Γ point, respectively. In contrast to Fe and Sn doping in TiCo0.85Fe0.15Sb that promotes band convergence,7 Zr and Hf might tend to induce the opposite effect. This trend suggests that the Ti orbitals mainly control the energy of the valence band edge at the Γ point, as evidenced in the ZrNiSn, NbCoSn and NbFeSb HH alloys by electronic band structure calculations.6 The isovalent substitution of Bi for Sb only weakly affects the α values at 300 K with values ranging between 78 and 96 μV K−1, indicating that Bi is either ineffective in inducing band convergence or that the highest concentration considered herein remains too low to observe any beneficial influence. At high temperatures, α tends to saturate, a behavior that can be ascribed to minority carrier excitation across the electronic band gap due to thermal energy.
These lowered α values are responsible for the decrease in the power factor PF = α2σ over the entire temperature range with respect to the values achieved in TiCo0.85Fe0.15Sb and in TiCo0.85Fe0.15Sb1−xSnx7 (Fig. 4(d)). Alloying with Bi has virtually no effect on PF at 300 K, while a marginal decrease is observed at higher temperatures.
The temperature dependence of the total thermal conductivity κ of each sample Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb1−xBix (0 ≤ x ≤ 0.04) is shown in Fig. 5(a). A sharp decrease in κ ∼ 62% at 300 K and ∼47% at ∼850 K is achieved upon Zr and Hf alloying with respect to TiCo0.85Fe0.15Sb. The addition of Bi further lowers κ by ∼29% to reach ∼2 W m−1 K−1 at ∼850 K in Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04. This substantial reduction in thermal conductivity is due to the enhanced disorder induced by these substitutions. As shown in Fig. 5(b), which presents a comparison with the current literature on both n- and p-type TiCoSb-based alloys, this value is among the lowest ever obtained in TiCoSb-based alloys.
Fig. 5 Temperature dependence of the (a) total thermal conductivity κ. (b) Comparison of κ of various TiCoSb-based HH alloys, with [A]-TiCoSb,45 [B]-Ti0.5Zr0.25Hf0.25Co0.95Ni0.05Sb,46 [C]-TiCo1.05Sb0.8Sn0.2,35 [D]-Ti0.6Hf0.4CoSb0.8Sn0.2,47 [E]-Ti0.5Hf0.5CoSb0.8Sn0.2,48 [F]-Ti0.9Nb0.1CoSb,49 [G]-Ti0.8Nb0.1Ta0.1CoSb0.8Sn0.2,50 [H]-Ti0.341Zr0.343Hf0.316Co0.957Pd0.043Sb,51 [I]-TiCoNi0.2Sb,52 [J]-Ti0.6Hf0.4Co0.87Ni0.13Sb,53 [K]-TiCoSb0.8Sn0.2,54 [L]-Ti0.6Zr0.4CoSb0.8Sn0.2,54 [M]-Ti0.5Hf0.4Ta0.1CoSb,34 [N]-TiCoSb0.9Sn0.1,55 [O]-TiFe0.3Co0.7Sb,56 [P]-TiCoGe0.2Sb0.8,57 [Q]-TiCo0.9Ni0.1Sb,58 [R]-Ti0.92Ta0.08CoSb,59 (c) Lorenz number L, electronic thermal conductivity κe and (d) lattice thermal conductivity κL. In panels (c) and (d), the horizontal dashed lines stand for the degenerate limit of L and the minimum lattice thermal conductivity κL,min, respectively. |
To further shed light on the influence of this heavy alloying on the heat transport, κ was decoupled into the electronic thermal conductivity κe and lattice thermal conductivity κL (Fig. 5(c) and (d)). κe was estimated according to the Wiedemann–Franz law κe = LσT, where L is the Lorenz number. L was calculated by the relation L = 1.5 + exp(−|α|/116) where L and α are expressed in 10−8 V2 K−2 and μV K−1, respectively.60 The L values are significantly lower than the degenerate limit L0 of 2.44 × 10−8 V2 K−2 and monotonically decrease with increasing temperature for each sample, as is frequently seen in heavily-doped semiconductors. Fig. 5(c) demonstrates that κe increases with temperature and does not vary significantly with the substitution of Bi for Sb. The lattice thermal conductivity κL = κ − κe decreases by ∼50% to reach ∼2.49 W m−1 K−1 at 843 K upon alloying on the Ti site. The κL is further reduced by Bi substitution, leading to the lowest κL value of ∼1.74 W m−1 K−1 achieved at 843 K in the sample Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04 (Fig. 5(d)). This lowest value is approaching the minimum limit κLmin = 0.907 W m−1 K−1, which was calculated for TiCoSb using the model developed by Cahill and Pohl61 (see equations in ESI†). The drop in κL over the whole temperature range primarily stems from the large enhancement in point-defect scattering caused by the substitutions of Zr, Hf, and Bi for Ti and Sb, respectively, which yield a pronounced increase in the mass and strain field fluctuations.
To further investigate the role played by these substitutions on κL, the Debye–Callaway model that depends on the relaxation time approximation and assuming the absence of dispersion in the vibrational spectrum with no variations between the longitudinal and transverse polarizations, was considered.62,63 The correction factor for phonon–phonon interactions is not taken into account in the simplified form of this model, which yields the following equation:63,64
Fig. 6(a) shows the disorder scattering parameters ΓM, ΓS, and Γ for all samples. All these parameters abruptly rise in the samples alloyed with Zr and Hf, indicating enhanced mass and strain field fluctuations in the crystal lattice due to the pronounced differences in atomic mass and size between Zr, Hf, and Ti. Moreover, a monotonic increase in ΓM, ΓS, and Γ is observed in the Bi-substituted samples, which favors further lowering in κL. The calculated value of Grüneisen parameter γ of each sample (Fig. S3 in the ESI†) shows a peak value of 3.4 for the alloy Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04, which is higher than that determined for TiCo0.85Fe0.15Sb. These high γ values are indicative of effective phonon–phonon scattering, which is a contributing factor to reduce κL.66
The relaxation time of various scattering mechanisms as a function of phonon frequency is depicted in Fig. S4 of the ESI.† The relaxation time curves for different scattering have been calculated for the representative sample Ti0.5Hf0.3Zr0.2Co0.85Fe0.15Sb0.96Bi0.04 at 450 K (higher than Debye temperature of TiCoSb), assuming that all the vibrational modes have been populated at this temperature and contribute to phonon scattering. High-frequency phonons are scattered primarily by point defects and Umklapp processes, while low-frequency phonons are scattered mostly by grain boundaries.62,64 These predictions have been proven in the current set of alloys, where it was shown that grain boundary scattering is important at low frequency but point defect and Umklapp phonon scattering dominate at high frequency. The point-defect scattering mechanism shows the minimal relaxation time of τPD ∼ 1 × 10−14 s when it is excited at the Debye frequency, indicating that this mechanism outweighs the other processes and is predominantly responsible for the suppression in κL observed experimentally. Fig. 6(b) shows the calculated κL for different combinations of scattering mechanisms, that is, (U), (U + PD) and (U + PD + B). The κL values measured at 450 K for all samples have also been added in Fig. 6(b). The combination of the three scattering mechanisms (U + PD + B) yields a good agreement, indicating that all these scattering mechanisms are influencing the heat transport in these alloys.
Fig. 7(a) shows the temperature dependence of the dimensionless thermoelectric figure of merit ZT. The co-substitution of Zr and Hf for Ti leads to a significant rise in the ZT values with a peak ZT of 0.27 at 843 K, which is ∼47% higher than that of parent alloy TiCo0.85Fe0.15Sb. The significant lowering in κL is the key factor behind this ZT improvement, which counterbalances the reduction in the power factor. The small concentration of Bi does not significantly impact the ZT values that reaches the moderate value of 0.3 at 843 K in the Ti0.5Hf0.3Zr0.2Fe0.15Co0.85Sb0.96Bi0.04 alloy. Finally, the engineering figure of merit (ZT)Eng was calculated as a function of the temperature difference ΔT within the cumulative temperature dependency (CTD) model (see ESI† for detailed information). (ZT)Eng increases with Zr, Hf and Bi doping, confirming that the reduction in thermal conductivity outweighs the decrease in the electronic properties. The peak (ZT)Eng value of ∼0.08 for ΔT = 500 K is realized for the x = 0.04 sample, which is ∼60% greater than that determined for TiCo0.85Fe0.15Sb as shown in Fig. 7(b).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ma00923h |
This journal is © The Royal Society of Chemistry 2023 |