Leah
Borgsmiller
and
G. Jeffrey
Snyder
*
Northwestern University, Department of Materials Science and Engineering, Evanston, IL 60208, USA. E-mail: jeff.snyder@northwestern.edu
First published on 9th September 2024
The Yb11Sb10 Zintl phase has long been reported to be a negative secondary phase in the high performing Yb14MSb11 (M = Mn, Mg, Zn) (14-1-11) thermoelectric material. However there have been discrepancies in the literature regarding the effect of this Yb11Sb10 phase on the thermoelectric performance of 14-1-11 compounds. The recent discovery of Yb10MnSb9 is used here to explain these inconsistencies. Yb11Sb10 and Yb10MnSb9 are very close materials stoichiometrically and crystallographically. However, despite these similarities, these materials exhibit very different transport properties and thus should have differing impacts on the overall thermoelectric performance of 14-1-11. In this work, we consider how these two phases affect the thermoelectric performance of Yb14MnSb11 when present as secondary phases both through an experimental investigation and an effective medium theory model. Secondary phases in thermoelectrics are often regarded as a consistently negative effect and yet here we show that while the presence of Yb11Sb10 in Yb14MnSb11 reduces the thermoelectric performance, the presence of Yb10MnSb9, even in large amounts does not dramatically reduce the thermoelectric performance and is shown here to even improve performance. This work provides a deeper understanding of previous works regarding 14-1-11 materials and Yb–Mn–Sb phases and paves the way towards practical application.
While there are a handful of material systems to date that have exhibited zTs greater than 1 at these high operating temperatures,9–12 making them candidates for practical applications, the search for new thermoelectric materials with promising properties is still on-going. The success of the high performing 14-1-11 Zintl thermoelectric compounds including Yb14MnSb11,13 Yb14MgSb11,14 and more recently Yb14ZnSb11,9 has lead to other discoveries of Zintl thermoelectrics with similar compositions, namely Yb21Mn4Sb1815 and Yb10MnSb9.16 Zintl phases are defined as being valence precise where the electropositive cations donate electrons to polyanionic covalent structures.17 Zintl phases are among some of the highest performing thermoelectric materials due to their often intrinsically low thermal conductivity from large complex unit cells and their semiconducting nature that together simultaneously allow for facile charge transport and impeded thermal transport.18–20
It can be seen that these Yb–Mn–Sb materials exist in a very busy region of phase space with multiple stable ternary and Yb–Sb binary compounds (Fig. 1), many of which can be described through Zintl–Klemm formalism. In particular, the Yb11Sb10 phase has been reported as an undesirable secondary phase that is often present, sometimes in large amounts, in 14-1-11 samples.29 This secondary phase is expected to harm performance as it has been reported to be a fairly metallic phase with a low Seebeck coefficient, high electrical conductivity, and high thermal conductivity.30 However, many reports indicate that samples with 10% or more Yb11Sb10 have transport properties largely unaffected by this metallic phase.29,31 This phase is also seemingly present, even when samples are made in Mn excess.29 This would not be expected if the sample was synthesized in thermodynamic equilibrium in a Mn-rich phase space given what is understood about this ternary phase diagram shown in Fig. 1. Additionally, in one particularly contrarian example in the thesis of Hu (2017),32 it was found that the presence of Yb11Sb10 in large amounts (∼50%) in Bi-doped Yb14MgSb11 actually improved the thermoelectric properties through an increase in the Seebeck coefficient and a decrease in the thermal conductivity. This observation cannot be explained solely by this Yb11Sb10 phase. It would not be expected this phase with an incredibly low Seebeck coefficient would contribute to increasing the Seebeck coefficient in 14-1-11 samples.
![]() | ||
Fig. 1 Schematic phase diagram showing the reported phases in the Yb–Mn–Sb phase space. Many of these phases are reported in computational databases like the Materials Project21 and OQMD,22,23 while others are found from experimental reports.13,15,16,24–27 The rectangle on the left is showing a zoomed-in region of the ternary diagram which is particularly busy around the high performing Yb14MnSb11 phase. The shaded three-phase region with Yb14MnSb11,13 Yb10.5MnSb9,16 and Yb11Sb10![]() |
However, we can reconcile these observations regarding Yb11Sb10 in 14-1-11 samples by considering the newly reported Yb10MSb9 material (M = Mn,33 Mg34). While the Yb11Sb10 phase has been known for decades,28 this Yb10MSb9 material was only first reported in 2021,16 years after much of the foundational work regarding Yb14MSb11 (M = Mn, Mg, Zn). Due to the incredibly similar crystal structures between Yb11Sb10 and Yb10MSb9, it is likely that some amount of Yb10MSb9 could be misidentified as Yb11Sb10.16
While our previous work has considered the thermoelectric properties of this new 10-1-9 phase through transport measurements,33,35 electronic structure evaluation,36 and doping studies,34 this work aims to tackle a different issue. Secondary phases have been frequently reported in these Yb–Mn–Sb Zintl phases and are difficult to control. Here we consider how the likely presence of this 10-1-9 phase impacts the thermoelectric performance of the Yb14MnSb11 material which has been of interest to the thermoelectric community for years.
In this work we will consider the implications of both Yb11Sb10 and Yb10MnSb9 as key secondary phases in Yb14MnSb11 through a combination of analytical and experimental approaches. To analyze the effect of these secondary phases, bulk polycrystalline Yb14MnSb11 samples were synthesized with purposeful addition of secondary phases Yb11Sb10 and Yb10MnSb9. The thermal and electronic transport of these samples as well as the X-ray diffraction (XRD) patterns were then compared to the pristine Yb14MnSb11 sample. Complementary computational techniques such as effective medium theory were then utilized to further explore the impacts of these phases on the thermoelectric performance of Yb14MnSb11. We show that while the presence of Yb11Sb10 alone is not sufficient to explain experimental trends with secondary phases in Yb14MnSb11, the additional consideration of Yb10MnSb9 as a secondary phase allows us to fully explain previously observed results. Additionally, the presence of Yb10MnSb9 in 14-1-11 samples is shown to even improve thermoelectric performance across certain temperature regimes.
In order to consider the most likely secondary phases, we must first consider the key sources of elemental loss in synthesis and subsequent characterization. While it is possible for various off stoichiometries and elemental loss to occur during synthesis of Yb14MnSb11 which is particularly likely for Mn,29 the largest source of elemental loss both in synthesis and subsequent experimental measurements is expected to be oxidation of Yb. Yb oxidizes quite readily, particularly at high temperatures, and this is a major problem for this material. In a study on the oxidation of 14-1-11 compounds, it was shown that large amounts of Yb2O3 form upon heating in air, particularly on the surface of the material.38 This results in a core material that is Yb-deficient and reportedly contains large amounts of the Yb11Sb10 phase.38 While high temperature thermoelectric synthesis and subsequent high temperature transport measurements are generally conducted under argon or high vacuum environments, any trace amounts of oxygen result in some Yb oxidation which has been observed through a discoloration of sample surfaces after measurement, even under high vacuum.
If we consider a model sample that starts out with a perfect 14-1-11 stoichiometry and then oxidizes, we can consider the expected resultant secondary phases. If this perfect composition sample only loses Yb through the formation of Yb2O3, we would expect both Yb11Sb10 and Yb10MnSb9 to be present as secondary phases in thermodynamic equilibrium. The respective phase fractions of these secondary phases can be predicted using simple ternary lever rule type approach, where the total composition must equal the sums of the respective compositions and concentrations of the constituent phases. This analysis is shown in Fig. 2. Here we can see that as Yb is removed from the bulk of the sample through surface oxidation, the phase fraction of Yb14MnSb11 continuously decreases with a corresponding rise in the amounts of Yb10MnSb9 and Yb11Sb10. This path in the three phase region is much closer to the Yb10MnSb9 phase than the Yb11Sb10 phase which can be seen graphically in the top left inset of Fig. 2 which shows this three phase region of interest with an arrow along the line of decreasing Yb. This means that in cases where the dominant source of elemental off-stoichiometry is Yb loss, the most present secondary phase would be expected to be Yb10MnSb9. In order for there to be more Yb11Sb10 than Yb10MnSb9 in a 14-1-11 sample, there would have to be substantial loss of Mn as well. This is certainly possible, especially depending on the synthesis conditions. Mn makes up a very small percentage of 14-1-11 by atomic ratio, which is even smaller by mass. Given that the primary sources of elemental loss are considered to be Yb followed by Mn, it is likely that most 14-1-11 samples are in this three phase region between Yb14MnSb11, Yb10MnSb9, and Yb11Sb10.
![]() | ||
Fig. 2 Expected phase fractions (by moles) of Yb14MnSb11, Yb10MnSb9, and Yb11Sb10 as a result of Yb loss through oxidation calculated through a ternary lever rule method. As more Yb is lost, the phase fraction of Yb14MnSb11 decreases whereas the amount of Yb10MnSb9 and Yb11Sb10 increases. The dashed line depicts the total amount of Yb10MnSb9 and Yb11Sb10 secondary phases. The triangle in the top left corner depicts this three phase region of interest from the phase diagram in Fig. 1, with the arrow depicting the direction of Yb loss. |
The key crystallographic similarities between Yb11Sb10 and Yb10MnSb9 are particularly apparent in the lattice parameters and the presence of Sb dimers amongst dispersed Yb cations. In the Yb10MnSb9 material there is also the presence of MnSb4 tetrahedron, similar to those seen in Yb14MnSb11. These crystallographic similarities, along with general unit cell complexity, complicates secondary phase identification.
Generally, when synthesizing thermoelectric samples, XRD is an early step used to determine the presence of secondary phases in the material. However, the more complicated the unit cell, the more complicated the XRD patterns which makes differentiation between these phases, particularly as secondary phases, non-trivial.
In the original study identifying the Yb10MnSb9 phase, Baranets et al. warn that the similar crystal structures between Yb10MnSb9 and Yb11Sb10 could result in misidentification.16 Further studies since then on the 10-1-9 structure have considered these structural similarities and differences.35,40 These structural analyses have concluded that distinguishing between these two phases using XRD should use key differentiating peaks in the 26°–29° 2θ range (for Cu Kα radiation). These peaks are shown in Fig. 3 and are indicated by the diamond markers in the Yb10MnSb9 pattern.
![]() | ||
Fig. 3 Comparison of expected XRD patterns (from CIF files generated from literature single crystal data) for Yb11Sb10,28 Yb10MnSb9,16 and Yb14MnSb11![]() |
However, while it is clear that we can use these peaks to differentiate between Yb10MnSb9 and Yb11Sb10, as has been done previously,35 these marked peaks overlap with higher intensity peaks in Yb14MnSb11. Thus, while XRD patterns of 14-1-11 samples can tell us if a sample contains some amount of Yb11Sb10/Yb10MnSb9, it is incredibly difficult to determine if those secondary phases are Yb11Sb10, Yb10MnSb9 or a combination of the two using XRD alone. Looking back at previous reports of Yb11Sb10 in Yb14MnSb11, it is quite plausible that at least some of these samples contain Yb10MnSb9 in addition to or instead of Yb11Sb10. Since using XRD alone is not a straightforward way to differentiate between these materials in 14-1-11 samples, a further investigation of the transport properties is beneficial in our understanding of these materials as secondary phases.
In order to experimentally investigate the impacts of these phases on Yb14MnSb11, three different samples were synthesized from the same batch of Yb14MnSb11 ball-milled powder. One sample was comprised of only this stoichiometric 14-1-11 powder, while the other two were synthesized with 20% by weight of Yb11Sb10 and Yb10MnSb9 respectively (see the methods section for more detail). This 20% secondary phases by weight corresponds to 17% secondary phases by molar ratio. This purposeful addition of secondary phases is reminiscent of the phase boundary mapping approach.7
Here we should note, that while we have added powder made with stoichiometric Yb11Sb10 and Yb10.5MnSb9, to stoichiometric Yb14MnSb11, the complexity of the phase diagram and XRD patterns makes it difficult to confidently synthesize phase pure samples. If we consider the previous discussion on the problem with elemental loss, it perhaps might be likely that any of these three “pure” stoichiometries contains some amount of the other two phases as secondary phases. Here the goal is to consider how these secondary phases are expected to impact performance and utilize purposeful addition of secondary phases to probe this. We aim to point to the complexity of the problem and show general trends with changing stoichiometries of these samples. There is expected to be some margin of error however on the amount of secondary phases in these materials. For instance, while we added 20% by weight of powder synthesized in the stoichiometry 10.5-1-9 to powder synthesized with the stoichiometry 14-1-11, the final material might not have exactly 20% by weight of the Yb10.5MnSb9 phase. If the starting 14-1-11 material became slightly Yb-deficient during synthesis, it might already have some undetectable amounts of the 10.5-1-9 phase, meaning that the final sample in this case could have more than 20% secondary phases. However, regardless of precise amounts of secondary phases, we will see in the following analysis that these two phases are expected to impact thermal and electrical transport when present in 14-1-11.
The XRD patterns for these three synthesized samples are shown in Fig. 4. In general, all three samples have peak patterns that match well to the theoretical pattern for Yb14MnSb11 shown in black. We can then consider if XRD can indicate to us that these samples contain the expected secondary phases of Yb11Sb10 and Yb10MnSb9 respectively. As discussed previously and shown in Fig. 3, the best way to differentiate between Yb10MnSb9 and Yb11Sb10 is to consider lower intensity low angle peaks in the pattern which we do here in Fig. 4b. It is clear that the samples containing the added secondary phases exhibit additional short peaks around 2θ values of 25.7° and 27.6° (indicated by stars in Fig. 4b). However both the sample with added Yb11Sb10 and the sample with added Yb10MnSb9 exhibit these additional peaks and thus it is difficult to visually detect differences between the two diffraction patterns. This can be attributed to the fact that the key peaks that are different between Yb11Sb10 and Yb10MnSb9 overlap with higher intensity 14-1-11 peaks as discussed with Fig. 3.
![]() | ||
Fig. 4 (a) XRD patterns for the three experimental Yb14MnSb11 samples (Yb14MnSb11, Yb14MnSb11 with 20 wt% Yb11Sb10, and Yb14MnSb11 with 20 wt% Yb10MnSb9 respectively), and the theoretical pattern for Yb14MnSb11 shown in black. (b) Zoomed-in section of the XRD patterns, the black stars are indicative of both Yb11Sb10 and Yb10MnSb9. Here it can be seen that while the samples containing extra phases exhibit additional low angle peaks, the key distinguishing peaks from Fig. 3 are not readily observed due to the overlap with 14-1-11 peaks. |
We then considered if Rietveld refinement41 could be utilized to differentiate between these samples. Here it was found that for the samples with purposeful addition of secondary phases, the refinement was improved when the secondary phases of Yb11Sb10 and Yb10MnSb9 were included in the refinement. However, differences in the final Rw values were not significantly different if the refinement considered just Yb11Sb10 and Yb14MnSb11versus just Yb10MnSb9 and Yb14MnSb11 for both samples. Thus, while Rietveld refinement and XRD as a whole are useful tools for determining if 14-1-11 samples have some Yb10MnSb9/Yb11Sb10 phases, it cannot concretely differentiate between these two secondary phases in small amounts, especially in these bulk polycrystalline samples of Yb14MnSb11. Additionally due to the nature of Rietveld refinement, it is always possible that in adding phases to the refinement, the improvements in fit are coming from the additional fitting parameters in a non-meaningful way. Perhaps higher resolution XRD on larger crystal samples could allow for better differentiation. However, much of the literature on 14-1-11 materials for thermoelectric applications use similar polycrystalline synthesis and XRD methods. Other common methods of secondary phase identification such as electron microscopy (SEM) and energy dispersive spectroscopy (EDS) that rely on compositional contrast are also difficult for this system due to the incredibly similar stoichiometries and thus minimal compositional contrast between phases.
![]() | ||
Fig. 5 Transport properties for single phase Yb14MnSb11, Yb11Sb10 and Yb10MnSb9 (a) electrical conductivity, σ, (b) total thermal conductivity, κ, and (c) Seebeck coefficient, α. |
The expected and observed effects of these secondary phases on transport can be considering through a dual approach utilizing both experimental data (Fig. 6) and effective medium theory (Fig. 7).
![]() | ||
Fig. 7 Effective medium theory predictions of transport for Yb14MnSb11 samples containing some amount of Yb10MnSb9 and/or Yb11Sb10. In each plot the black line corresponds to experimental data for phase pure Yb14MnSb11 (a) the predicted effective electrical resistivity for samples of Yb14MnSb11 containing 17%, 25%, and 50% (mole%) of Yb10MnSb9 and Yb11Sb10 respectively. Note that this does not match experimental trends for the sample containing Yb11Sb10, pointing to the possibility of interfacial effects. (b) The predicted effective Seebeck coefficients for samples of Yb14MnSb11 containing 17%, 25%, and 50% (mole%) of Yb10MnSb9 and Yb11Sb10 respectively. (c) Predicted Seebeck coefficient curves for samples that are Yb deficient and have secondary phase fractions as predicted from Fig. 2. Here samples that range between 100% Yb14MnSb11 and 50% Yb14MnSb11 have very minimal differences in the Seebeck curve. |
We can next consider the impact of secondary phases on the Seebeck coefficient. Here in Fig. 6b, it can be seen that the presence of Yb10MnSb9 slightly increases the Seebeck coefficient (until very high temperatures where Yb10MnSb9 exhibits bipolar conduction before Yb14MnSb11). Conversely, the presence of Yb11Sb10 reduces the Seebeck coefficient as one might expect from the addition of a more metallic phase. We can consider the Seebeck coefficient to be a direct reflection of the carrier concentration of a material due to the relationship between α and Fermi level. Thus, we would expect a more metallic sample with a lower Seebeck coefficient (the one with additional Yb11Sb10) to have a higher carrier concentration, whereas a more insulating sample with a higher Seebeck coefficient should have a lower carrier concentration (the sample with the additional Yb10MnSb9). However, the differences in Seebeck coefficient and thus overall carrier concentration between the samples synthesized here are relatively small, due to the fact that these samples with additional secondary phases are still primarily 14-1-11 which has a relatively high undoped carrier concentration. Hall effect measurements of the samples measured at NASA JPL are provided in the ESI, Fig. S1.† All samples have a carrier concentration around 1 × 1021 cm−3 with the noise in the measurement greater than the differences between samples.
Note that in this work no dopants are used because Yb14MSb11 (M = Mn, Mg, Zn) achieves a maximum zT > 1 without the use of any dopants. This occurs because the compound has a valence imbalance leading to 1 hole per formula unit. The other semiconducting phase, 10-1-9 is also intrinsically doped, meaning that there are intrinsic defects, typically cation vacancies, that lead to excess holes. The three phase equilibrium between 14-1-11, 11-10 and 10-1-9 ensures that the intrinsic defect concentration is the same for any compositional mixture of the three phases.7
In this study we also report the trends in total thermal conductivity (Fig. 6c) and zT (Fig. 6d) with temperature and these secondary phases. It can be seen that the presence of either secondary phase reduces the thermal conductivity as expected from the corresponding increase in electrical resistivity. Calculating the zT using the measured transport data, shows that the presence of Yb10MnSb9 increases the zT slightly across this temperature range, whereas the presence of Yb11Sb10 decreases the zT slightly. However it should be noted that at higher temperatures where Yb14MnSb11 is of most interest, we would expect the zT of pure Yb14MnSb11 to be greater than that of Yb14MnSb11 with Yb10MnSb9 due to the earlier bend over in the Seebeck coefficient with Yb10MnSb9.
A commonly used formulation of the electrical conductivity using effective medium theory is as follows:46
![]() | (1) |
We can use this equation to consider what trends we might expect in the resistivity when Yb11Sb10 and Yb10MnSb9 are present in Yb14MnSb11. This is shown in Fig. 7a, where there is experimental data for Yb14MnSb11, and then predicted effective medium theory resistivities for Yb14MnSb11 with 50% (mole percent) Yb10MnSb9, 25% Yb10MnSb9, 17% Yb10MnSb9, 17% Yb11Sb10, 20% Yb11Sb10, and 50% Yb11Sb10. Because Yb11Sb10 is metallic with a very low electrical resistivity (see Fig. 5), we would expect the presence of Yb11Sb10 to decrease the resistivity in Yb14MnSb11. Conversely, Yb10MnSb9 is much more electrical resistive and would be expected to increase the resistivity of Yb14MnSb11. However in comparing this result to the experimental result of Fig. 6 it can be seen that it was observed that both the presence of Yb11Sb10 and Yb10MnSb9 increased the overall sample resistivity. Thus while this formulation for effective medium theory conductivity held well for Yb11Sb10–Yb10MnSb9 composites,35 this doesn't hold for these samples. This could be due to a number of reasons. To start, this effective medium theory formulation is quite simple and assumes spherical and isotropic phases. It also assumes that there is no effect from the interfaces between phases. In real materials, we know that grain boundaries can have significant impact on transport, which could be occurring here. Additionally, these samples were synthesized by hotpressing powders of Yb14MnSb11 and crushed secondary phases pellets which could further exacerbate this interfacial effect. Thus, here using the dual approach of effective medium theory and experimental results, we can conclude that there are likely interfacial effects that are impacting the electrical resistivity. While uncovering the exact role of these interfaces would be difficult experimentally due to differing synthesis routes and difficulties in imaging the microstructure of these materials, we believe this would be a fascinating area of potential future work.
As was done with the electrical resistivity, we can consider the expected impact on the Seebeck coefficient through an effective medium theory approach. While the electrical conductivity can be largely affected by grain boundaries and interfacial effects, the Seebeck coefficient is expected to be relatively unaffected. We can use the following expression,49 where χi is the phase fraction of the specific phase, i, σi is the electrical conductivity of the phase, and αi is the Seebeck coefficient of the phase. The terms σeff and αeff represent the overall expected effective values.
![]() | (2) |
In this work we utilize this expression because it can be more easily extended to a three-phase system unlike other expressions that are only considered for the simplest two-phase system.48 Once again, this simple formulation of effective medium theory assumes that all secondary phase regions can be approximated to be spherical and isotropic. This formulation has been also shown to be most accurate for systems involving degenerate semiconductors and metals.49 Additionally, it should be noted that this equation requires inputs of not only the Seebeck coefficients for the constituent phases but also the electrical conductivities of constituent phases and the overall effective conductivity. Here we utilize the effective conductivity equation (eqn (1)), although we have already shown that while this seems to describe the effect of Yb10MnSb9 well, it doesn't well describe the effect of Yb11Sb10, perhaps due to additional electrical interface resistance between Yb14MnSb11 and Yb11Sb10. Thus while the overall trends (increasing/decreasing) can be considered via this simple model, the predicted values may not be as accurate, particularly for the case of Yb11Sb10.
Considering now the predicted Seebeck curves in Fig. 7b, we can see that the Seebeck is expected to increase with the addition of Yb10MnSb9 and decrease with the addition of Yb11Sb10 which qualitatively agrees well with experimental results in Fig. 6b. It should also be noted that while the sample labeled Yb14MnSb11 was synthesized using a pure stoichiometric ratio without any purposeful addition of secondary phases, based on this analysis we believe it is plausible if not likely that there are still trace amounts of secondary phases formed such as through Yb loss from oxidation. If this were the case, there could be additional Yb10MnSb9 in all three samples beyond what was purposefully added, meaning that we might expect there to experimentally be a larger increase in Seebeck for the sample with added Yb10MnSb9 and a smaller decrease in Seebeck for the sample with added Yb11Sb10 which is seen here.
We can also use the phase fraction curves in Fig. 2 to model the expected Seebeck coefficients upon Yb loss through oxidation which is modeled in Fig. 7c. Here we show the experimental Seebeck curve for Yb14MnSb11, as well as curves for modeled samples with 90%, 80%, and 50% Yb14MnSb11, with the remaining 10–50% secondary phases that are a combination of Yb10MnSb9 and Yb11Sb10 as predicted from Fig. 2.
Here it can be seen that three-phase samples with 0–50% secondary phases, exhibit very little differences in the overall Seebeck curve. These differences are on the order of magnitude of potential error you might expect from a Seebeck experimental measurement. In contrast to the results in Fig. 7b which show that samples with 50% Yb10MnSb9 should demonstrate a more significant increase in Seebeck coefficient, and samples with 50% Yb11Sb10 should demonstrate an even more dramatic decrease in Seebeck coefficient, samples with 50% secondary phases containing a mixture of Yb10MnSb9 and Yb11Sb10 only change the overall Seebeck coefficient slightly.
If we consider a different study in the thesis of Hu (2017),32 it was reported that samples of Bi-doped Yb14MgSb11 with 50% of “Yb11Sb10” exhibited the highest zT, due to an increase in the Seebeck coefficient, increase in the electrical resistivity and decrease in the thermal conductivity compared to samples with less (but still present) secondary phases. This trend cannot be explained by the presence of Yb11Sb10, however it falls perfectly in alignment with what would be expected for a sample with substantial 10-1-9 phase that has been misidentified as Yb11Sb10. Note that in this study they were considering the Mg-analog of 14-1-11, Yb14MgSb11. However both Yb14MnSb11 and Yb14MgSb11 have incredibly similar transport9 and so do Yb10MnSb9 and Yb10MgSb9.34 Thus we would expect the trends of 10-1-9 in 14-1-11 to be roughly the same for the Mg case as is studied here for the Mn case. In consideration of this study, we show that the presence of Yb10MgSb9 as a secondary phase can explain previously contradictory results.
In addition to using this analysis to explain past results, we can use these results to predict the potential evolution of performance of a sample with time. A key consideration in designing thermoelectrics for practical applications is the stability of the material and how it evolves on timescales of utilization. In the case of Yb14MnSb11, the key evolution we might expect with time would be oxidation, with higher oxidation rates expected under conditions with higher oxygen environments and higher temperatures. If Yb from the sample is loss through oxidation, we would expect the breakdown of constituent phases to follow the trend discussed previously in Fig. 2. This means that with time we would expect the sample to continually decrease the amount of Yb14MnSb11 while substantially increasing the amount of Yb10MnSb9 and slightly increasing the amount of Yb11Sb10. This would correspondingly result in an observed shift in the XRD patterns indicating large amounts of these secondary phases. However, as we've shown here, Yb10MnSb9 is not expected to substantially harm the thermoelectric performance as we would expect Yb11Sb10 to do, and it has the potential to even improve performance. Thus, despite expected and observed changes in the phase purity with cycling, the thermoelectric performance is not expected to be dramatically harmed and thus this evolution is not as important of an issue as previously suspected for long term performance.
First the Mn was pre-reacted with Sb forming a binary precursor MnSb as has been done in other 14-1-11 studies to mitigate the formation of undesirable Yb–Sb binaries.29,37 This was done by combining Mn and Sb in a 1:
1 molar ratio and sealing in a quartz tube under vacuum. This tube was then heated to 950 °C in a furnace at a rate of 100 °C per hour and held at temperature for 2 hours. Then the temperature was decreased to 675 °C and held there for 48 hours before an air quench by removing the tube from the furnace. The complete reaction was confirmed using powder XRD on a STOE STADI P instrument using Cu Kα radiation.
In order to synthesize the Yb11Sb10 and Yb10.5MnSb9 samples, the Yb, Sb and MnSb (in the case of Yb10MnSb9) were measured in their respective atomic ratios totalling 7 grams each. These components were loaded into a stainless steel ball mill jar containing stainless steel balls in an argon glovebox and sealed with an O-ring. These mixtures were then mechanically alloyed through high energy ball-milling in a SPEX high energy ball mill. This was done for three 90 minute intervals, with the sides of the jar being scraped with steel tools in between each cycle to aid in the reaction.
This ball-milled powder was then hot-pressed into individual 1/2 inch diameter pellets using a home-built induction hot press51 and graphite dies with graphoil and graphite spacers. These samples were pressed using ∼2.3 g of powder each and pressed at 800 °C for 45 min at an average pressure ∼45 MPa. XRD data for these as synthesized stoichiometric Yb11Sb10 and Yb10.5MnSb9 pellets are supplied in the ESI, Fig. S2.†
For the 14-1-11 samples, a single batch of powder with a synthesis composition of Yb14MnSb11 was made by combining Yb, MnSb, and Sb and ball-milling as described for the samples above totalling 10 grams. From this single batch of powder, three different samples were synthesized. The first sample contained only this Yb14MnSb11 powder totalling approximately 2.5 g. The second sample contained 2 g of this Yb14MnSb11 powder and 0.5 g of crushed Yb11Sb10 pellet (20% by weight) that was crushed using a mortar and pestle. Similarly, the last sample contained 2 g of Yb14MnSb11 and 0.5 g of crushed Yb10.5MnSb9 pellet. All three pellets were then hotpressed using the same methods as discussed for the previous series of samples. A second set of identical samples were then made and used for the XRD and high temperature transport measurements conducted at NASA Jet Propulsion Laboratory (JPL).
Higher temperature electrical transport measurements as shown by the solid lines in Fig. 6a and b, were measured at the NASA Jet Propulsion Laboratory using similar instruments up to 1000 °C.
Thermal transport was considered by measuring the thermal diffusivity using a Netzsch LFA 457 MicroFlash instrument. This measured thermal diffusivity was then converted to a thermal conductivity using the Dulong–Petit heat capacity and sample density measured using the Archemides method in isopropyl alcohol.
The electrical resistivity and thermal conductivity measurements done here (indicated by the markers in Fig. 6a and c) were measured and reported for heating curves. The Seebeck measurements done here (markers in Fig. 6b) as well as the Seebeck and resistivity curves measured at JPL (solid lines in Fig. 6a and b) were conducted in both heating and cooling conditions and the data points shown here are averages between the two curves rather than explicitly measured data points. The full data set is supplied in the ESI.† While there was some hysteresis, it was considered to be minimal and does not change the conclusions of this work.
X-ray diffraction measurements were conducted on bulk polycrystalline pellets in reflection mode on a STOE STADI MP instrument using Cu Kα radiation. Rietveld refinement was conducted using the GSAS-II software.54
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04606d |
This journal is © The Royal Society of Chemistry 2024 |