1. Introduction
The total amount of energy produced nowadays on a global scale that is lost as waste heat reaches up to a remarkable amount of around 66% [
1]. Using this waste heat as a power source is tremendously beneficial. Thermoelectric (TE) devices comprise materials that can convert waste heat into useful electricity and vice versa, based on the Seebeck and Peltier effects. These devices can be rather reliable and compact as they do not contain moving parts and can be used in harsh environments, making them suitable for a variety of applications, including space ones [
2].
TE efficiency is best defined by the figure of merit, which is equal to the following:
where
σ is electrical conductivity;
S is the Seebeck coefficient;
T is the absolute temperature; and
κ =
κΙ +
κe is the total thermal conductivity, comprising the lattice (
κΙ) and electronic (
κe) counterparts. Ways to increase
zT include maximizing the Power Factor (PF),
σS2, while minimizing thermal conductivity. Recent advances in this remedy focus on band reforming and nanostructuring approaches, such as the introduction of grain boundaries, point, line and planar defects and elemental alloying [
3]. The ultimate goal is to considerably suppress
κ, while preserving a high PF and, hence, a significantly high
zT.
Full or half-Heusler (HH) materials have been widely known since the beginning of the 20th century due to their ferromagnetic behavior and, in recent years, for their promising spintronics applications [
4]. In addition, they exhibit promising TE properties for high temperature applications due to their enhanced power factor values and prolonged stability at higher temperatures [
5]. However, they suffer from relatively high thermal conductivity [
6], which lowers their performance [
7]. Ways to accommodate this include alloying and nanostructuring approaches [
6], with a view to increase phonon scattering through alloy elements fluctuations, imposed strain and the creation of a significant density of grain boundaries.
Among the various n-type HHs, ZrNiSn or HfNiSn and their doped counterparts exhibit improved TE performance as a result of their nanostructural features. Studies by Chauhan et al. [
3] and Du et al. [
8] pointed out the necessity of introducing lattice point defects, such as interstitials or substitutional ions, as an effective way of scattering short-wavelength phonons and hence reducing lattice thermal conductivity. However, the exact nature, type, locations and distribution of such defect centers are yet to be clarified to effectively design n-type HHs with enhanced TE properties, such as low thermal conductivity, and a high Seebeck coefficient, power factor and
zT.
Recent structural analyses employing electron microscopy are focused on systems for renewable technologies and energy harvesting, with TE materials being among them [
9]. Transmission electron microscopy (TEM) and associated techniques have become indispensable tools for unraveling the intricate nanostructure of HH materials, particularly those based on the multicomponent Ti-Zr-Hf-Ni-Sn-Sb system. These complex alloys exhibit a rich microstructure influenced by multi-element substitution, which can significantly impact thermoelectric performance. In a study by Morimura et al. [
10], high-resolution TEM (HRTEM) was employed to investigate the structural features of Ti
0.5(Zr
0.5Hf
0.5)
0.5NiSn
0.998Sb
0.002 alloys. The authors observed nanoscale domain structures, typically ranging from 2 to 10 nm in size, that revealed varying degrees of atomic ordering. Fourier and inverse Fourier transformation analyses confirmed the coexistence of both ordered and disordered phases, particularly in the second nearest-neighbor positions, suggesting a complex local atomic arrangement that contributes to phonon scattering and affects thermal conductivity. In another investigation, Populoh et al. [
11] examined multiphase Ti-Zr-Hf-Ni-Sn systems and reported phase separation into Ti–rich and Zr/Hf–rich domains at a nanometer scale. TEM observations confirmed the coexistence of different half-Heusler phases while energy dispersive X-ray spectroscopy (EDS) revealed that Ni and Sn remained uniformly distributed, whereas the elemental segregation of Ti, Zr and Hf led to compositionally distinct nanodomains. This phase separation was linked to enhanced phonon scattering and improved thermoelectric performance, with
zT values reaching up to 1.0 in a temperature range from 675 to 775 K.
The nanostructural features of binary and ternary HH materials, of the general type (Ti
0.4Zr
0.6)
1−xHf
xNiSn (x = 0, 0.3), are dealt with in a comparative manner in the current study. The selected HH materials, doped with low amounts of Sb (1.5 and 2 at%, respectively), were synthesized by an easy, inexpensive and timely one-step mechanical alloying method [
12,
13], and they have shown a remarkable improvement in their TE properties, especially power factor, electrical conductivity and resulting figure of merit
zT. Variations in particle size, morphological characteristics, phase multiplicity, crystalline quality and lattice constant variations have been assessed by a combination of TEM/HRTEM, selected area diffraction (SAD), EDS and post-experimental HRTEM image analysis employing Geometrical Phase Analysis (GPA) [
14]. Structural features mainly responsible for TE property enhancement form the basis for synthesizing tailor-made TE nanomaterials with considerable alloying, i.e., high entropy TEs, while preserving optimum TE performance by not sacrificing mobility, thermal or electrical conductivities.
3. Results and Discussion
HH compounds crystallize in the MgAgAs-type structure at the high symmetry F
3m space group (#216), comprising three interpenetrating fcc sublattices. In principle, in an ideal unit cell, Sn ions occupy the octahedral 4a (0, 0, 0) sites (C sites) and Ti/Zr/Hf ions occupy the 4b (½, ½, ½) sites (N sites), whereas Ni ions occupy one of the two tetrahedral 4c sites, (¼, ¼, ¼) [or, equivalent (¼, ¼, ¾)] (T sites) [
15].
Mechanical alloying, followed by hot press sintering of the samples resulted in the formation of a single HH phase, both for the binary (Ti-Zr) and ternary (Ti-Hf-Zr) materials. This has been previously identified by X-ray diffraction [
12,
13], where the formed HH phase in each case exhibited narrow XRD peaks, after a minimum milling time of 7 h at a speed of 600 rpm. In addition, the binary Ti
0.4Zr
0.6NiSn
0.985Sb
0.015 sample showed the presence of a Ni
3Sn
4 phase, even after hot pressing; such impurities are a common feature in this system. Interestingly, this impurity did not exist in the Hf–containing samples.
The morphology and particle size distributions of the HH phase in the samples are depicted in the TEM images of
Figure 1. The HH particles appear agglomerated, with no specific shape and have an average size of 115 nm in Ti
0.4Zr
0.6NiSn
0.85Sb
0.015 and 223 nm in (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02. Their size distributions range from 60 to 210 nm and 95 to 405 nm, respectively; hence, Hf incorporation brought upon a significant increase in HH particle sizes. Nevertheless, the average grain size of our samples remains low, compared to the ~300 nm reported so far [
16]. The corresponding SAD patterns in
Figure 1a,b reveal that they are single crystalline materials; the [011] zone axis pattern of
Figure 1b particularly supports this. Measurements of the reciprocal lattice spacings assigned them to mixed HH phases; the results are listed in
Table 1.
The theoretical values of the lattice constants were calculated from the HH end members TiNiSn (
aTiNiSn = 0.594 nm), ZrNiSn (
aZrNiSn = 0.612 nm) and HfNiSn (
aHfNiSn = 0.609 nm) using Vegard’s law. The experimental values, both from XRD [
12,
13] and SAD, comply well with the theoretical ones. In addition, Zr and/or Hf substitution for Ti in the binary or ternary HH compounds resulted in an increase in the lattice constant; the TiNiSn end member possesses by far the smallest lattice constant among the three of them.
As a consequence of ball milling synthesis, the samples contain a significant number of nanoparticles and nanograins, agglomerated with larger particles. Nanograin morphology is illustrated in the two complementary bright- (BF) and dark-field (DF) images of
Figure 2a,b, respectively, from the ternary (Hf–containing) sample. The nanograins are clearly distinguished in the DF image. Along with the large particles, they also exhibit high crystallinity, as indicated by the Moiré fringe patterns created in several areas of the BF/DF images (circled in the DF image for clarity), as a consequence of their superposition on top of them or on the large particles. The coexistence of numerous large and, especially, small nanoparticles in patchy morphology and dense agglomeration in the HH samples typically results in the creation of a large proportion of grain boundaries that effectively act as scattering centers for phonons, leading thus to a reduction in thermal conductivity, as already manifested, initiated from binary Ti
0.4Zr
0.6NiSn
0.85Sb
0.015.
Nanoparticle formation has been best proven by HRTEM experiments.
Figure 2c presents agglomerated nanoparticles in Ti
0.4Zr
0.6NiSn
0.85Sb
0.015 whereas a discrete morphology is depicted in
Figure 2d from (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02. Measurements of their size distributions from HRTEM images resulted in 5–15 nm for Ti
0.4Zr
0.6NiSn
0.85Sb
0.015 and 4–17 nm for (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02. Although a greater size and wider distribution has been measured for the large particles in ternary HH, both samples exhibited similar ranges and sizes for their small nanoparticles. This can be attributed to ball milling synthesis of both materials, which produces particles with similar size distributions at the nanoscale level. The high percentage of small nanoparticles in ternary (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 compensated for the greater size of the large particles in it. The nanoparticles in both HRTEM images mainly exhibit (200) lattice fringes, whereas the single crystalline character is also manifested by the Moiré fringes there. Detailed measurements of the lattice fringe spacings in similar HRTEM images of both samples have been performed in order to derive the ‘nanometer’ scale lattice constant (
aHRTEM), and the results are included in
Table 1. There, it is shown that the average lattice constants in the nanoscale are in good agreement with the theoretical ones for both samples. However, it should be noted that the
aHRTEM constant range for the Hf–containing sample was a bit greater, ranging from 0.587 nm to 0.628 nm, taking into account an error of 0.5% in the measurements. This indicates that, although the experimentally measured average lattice constant of the HH nanoparticles appears to approach the theoretical one, there are certain deviations, which imply elemental fluctuations, point defects, interstitials and atomic substitutions in the mixed HH phases. This is in very good agreement with an EXAFS analysis previously published for samples belonging to the same family of HH materials [
15].
Energy dispersive X-ray spectroscopy (EDS) confirmed the aforementioned findings.
Figure 3 shows two representative EDS spectra from the binary and ternary HH materials. The spectra show all element peaks of the compounds, as well as the dopant (Sb). The latter, however, was difficult to distinguish due to its low content (1.5 at% and 2 at%, respectively) and overlap with multiple Sn L-family lines; therefore, the combination of these two elements is considered in any quantitative calculations from now on. EDS spectra did not reveal any other impurities; interestingly, O content was also negligible in the particles, which illustrates the chemical stability of the ball-milled TE materials [
12].
Quantitative analysis showed that, on average, the elemental ratios are in good agreement with their theoretical contents for both materials, taking into account an EDS accuracy of 3–5% and provided that the sums of the elements in their nominal sites (Ti + Zr + Hf) and (Sn + Sb) are predominately considered. However, there are elemental variations among distinct particles, which tend to be more pronounced in the ternary (Hf–containing) sample. The results are presented in
Table 2 in more detail and show that the elemental variations are more pronounced for Zr and Hf contents in particular, which implies elemental fluctuations among the HH particles down to the nanometer scale. This is in line with other studies [
11] where Ti–, Hf– and Zr–based nanoscale domains co-existed within a HH matrix, with a positive effect on TE performance. On the other hand, Ti content tends to be more stable in our materials. To further support this interesting finding, regions in (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 have been detected with increased Hf content.
Figure 4 confirms the latter case, where a particle with an increased Hf content domain is shown, along with its EDS spectrum inset. Hf content in these particle regions is three times higher than the nominal content, whereas there are greater deficiencies in Zr and Ti contents (−64 and −51 at%, respectively) compared to Ni and (Sn + Sb) ones; −16 and −12 at%, respectively, for these two crystallographic site elements. This reveals that Hf predominately substitutes for its corresponding N site elements in such Hf–rich particles [
15]. Having clarified this, any remaining Hf content was found in either Hf–rich particle locations or, potentially, other crystallographic sites in the HH lattice. This is in line with TE property measurements [
13], where a decrease in the Seebeck coefficient and an increase in carrier concentration have been reported as a consequence of Hf incorporation. Furthermore, as the substitution of Zr by Hf is isoelectronic, the increase in carriers has been attributed to defect creation (vacancies and/or interstitials), which is confirmed by ion substitution, vacancies or interstitial phenomena (i.e., alloy fluctuation) as implied by SAD/HRTEM and shown by EDS.
It should be clarified that the Hf–rich particle in
Figure 4 represents a minority feature of the sample, merely identified due to the resolving power of TEM. As a more representative outcome, less Hf–rich regions have been sometimes encountered, where Hf excess was about twice as high, whereas Zr, Ti, Ni and (Sn + Sb) deficiencies reached −15, +1 (i.e., excess), −13 and −9 at%, respectively. This shows that Hf incorporation in the ternary HH has a diverse substitutional behavior, ranging from Hf–rich domains in discrete particles with excessive substitution for Zr and Ti, to particles with concentrations closer to the nominal case. In both cases, however, certain deviations from stoichiometry for the other crystallographic site elements (Sn/Sb for C sites and Ni for T sites) are frequently observed to the same extent (10–15 at%). Beyond this, any variations in Sn content in both doped sample categories may be attributed to the peak overlapping with Sb, which shows a very high content for the same reason. The total (Sn + Sb) content is more accurate in this case, which indeed shows a small difference compared with the nominal value.
TEM characterization also sheds light on the existence of additional minority phases often encountered in HH TEs [
12]. In particular, the presence of discrete Ni
3Sn
4 particles has been confirmed by SAD and EDS measurements only in the binary HH sample. This distinct morphology is best presented in
Figure 5a, where the black-arrowed Ni
3Sn
4 particle has a size of ~120 nm and is crystalline as proven by its SAD pattern (inset). The EDS spectrum in
Figure 5b further confirms this, whereby the Ni and Sn peaks are far more prominent in relation to the Zr, Ti and Hf ones, the latter three stemming from overlapping HH particles. Taking into account that TEM can easily identify discrete phases at the nanoscale, it has to be stated that such phases were seldom encountered in our materials; the absence of such additional phase formation in (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 further proves this.
Lattice parameter variations can affect TE properties, thermal conductivity in particular [
17,
18]. These may stem from deviations from the nominal chemical composition, attributed to substitution ions, point defects or interstitials. In order to assess this contribution, the lattice constant was probed by Geometrical Phase Analysis (GPA) [
14] on HRTEM images, and characteristic results are presented in
Figure 6 for both large and small particles in the ternary (Hf–containing) material. GPA analysis from a large (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 particle (
Figure 6a,c) using the
g111 spatial frequency reveals a gradual transition of the lattice constant reaching up to 0.8% along the area indicated in
Figure 6c (spatial resolution 1.4 nm and standard deviation 0.5%). On the other hand, in smaller areas such variations are not discernible as shown for a small (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 nanoparticle presented in
Figure 6b and the respective GPA analysis in
Figure 6d, showing variation in the
g200 spatial frequency. A uniform distribution of the lattice constant is demonstrated with a 0.4% standard deviation using 1.2 nm spatial resolution. Overall Hf incorporation in the matrix may result in lattice constant variations over larger areas due to elemental fluctuations as evidenced by the TEM methods used in this study.
The nanostructural features revealed by the combined TEM methods had a significant influence on TE properties; the latter were presented in previously published works [
12,
13]. There, both Ti
0.4Zr
0.6NiSn
0.85Sb
0.015 and (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 showed lower thermal conductivity values; prior to Sb doping, −2.63 W/mK, and 2.1 W/mK at room temperature, respectively. Such values are lower, in general, than the ones reported for samples synthesized by arc melting [
2]. The reduced particle sizes along with the considerable elemental fluctuations and resulting lattice constant variations revealed in this TEM study resulted in such low
κ values. Even for the ternary, Hf–containing sample, the larger particle size observed by conventional TEM imaging was compensated for by the increased elemental fluctuations caused by Hf incorporation as deduced by SAD, HRTEM and EDS. In addition, the latter sample showed a significant amount of nanostructuring, preserving the low range of size values, as in Ti
0.4Zr
0.6NiSn
0.85Sb
0.015. Concerning electrical properties, both the binary and ternary samples exhibited (prior to Sb doping) a high (absolute) Seebeck coefficient of about −205 μV/K and very promising power factors in the range of 22–26 μW/cmK
2 at 300 K; in addition, the ternary, Hf–containing sample preserved a good electrical conductivity of ~220 S/cm, comparable to 205 S/cm for the binary sample. Despite the fact that Hf substitution for Zr is isoelectronic, the increased carrier concentration that led to high
S and
σ [
19] can be attributed to a significant density of Hf–driven defects that act as donors. This is in full agreement with Hf–induced alloy fluctuations identified by the TEM analysis. Sb doping predominately brought upon a significant increase in electrical conductivity in both samples, −1681 S/cm and 2250 S/cm at 300 K for the binary and ternary samples, respectively, revealing a donor concentration increase by the dopant. This led to a significant increase in power factors, having a maxima of 32 μW/cmK
2 for Ti
0.4Zr
0.6NiSn
0.85Sb
0.015 and 38 μW/cmK
2 for (Ti
0.4Zr
0.6)
0.7Hf
0.3NiSn
0.98Sb
0.02 at 300 K. The increased power factors and relatively low thermal conductivities of the samples resulted in high
zT values of 0.71 and 0.76 at 800 K and 762 K, respectively.