Thomas A. Seymour-Cozzini,
Rae A. Earnest‡
and
Julia V. Zaikina
*
Department of Chemistry, Iowa State University, Ames, Iowa 50011, USA. E-mail: yzaikina@iastate.edu
First published on 21st July 2025
Reactive metal hydrides, SrH2 and BaH2, were used to make the novel thermoelectric skutterudite Sr0.92Fe2.98±xNi1.00±xSb12 and the previously studied skutterudite Ba0.9Fe3.00±xNi1.00±xSb12, demonstrating the advantages of the diffusion-enhanced hydride synthesis. High-temperature in situ powder X-ray diffraction showed that BaH2, Fe, Ni, and Sb react to form binary and pseudo-binary intermediate products, which further react with one another and the remaining Sb to form BaFe3NiSb12 in a narrow temperature range around 855 K, the optimal synthesis temperature of BaFe3NiSb12. Statistical analysis, the fundamental basis of design of experiments (DoE) methodology, was tested using 12 synthesis attempts, and indicated 861 K was the optimal annealing temperature for BaFe3NiSb12 synthesis, validating the utility of statistical methods for the optimization of synthesis conditions in multicomponent solid-state systems. Synthesis optimization for SrFe3NiSb12 was performed entirely using DoE tools, demonstrating the efficacy of DoE to narrow down the set of conditions needed to form single-phase products using the fast hydride route. Synthetic insensitivity to the ratio of Fe to Ni enabled tuning of electronic transport properties, resulting in a peak thermoelectric figure of merit (zT) of 0.54 at 673 K for Sr0.92Fe3.28Ni0.70Sb12.
The heat conversion efficiency of a thermoelectric material increases with an increasing figure of merit, zT, which is calculated based on the absolute temperature, T, Seebeck coefficient, S, thermal conductivity, κ, and electrical resistivity, ρ, according to the equation. Therefore, the ideal thermoelectric material has low thermal conductivity, low electrical resistivity, and a high Seebeck coefficient, maximizing zT.4,5 The Seebeck coefficient, electrical resistivity, and thermal conductivity are all related to the charge carrier concentration in a given material. This interdependence makes the optimization of zT difficult. Recent leaps in materials design have seen zT double6 over previous state-of-the-art materials, increasing from 1 to approximately 2. This is much closer to the zT of 3 expected to make thermoelectric materials a commercially viable means of electrical power generation.7 Advances in caged materials made since 1995 largely stem from Slack's phonon-glass electron-crystal concept,8 which decouples lattice thermal conductivity from the electrical properties of a thermoelectric material. A phonon-glass electron-crystal is a hypothetical material that can scatter thermal phonons effectively as if it is amorphous (glass) but can also conduct electricity well as if it is crystalline.8 Attempts to realize this decoupling of electrical and thermal properties are most successful in compounds with cage motifs in their crystal structure. Filled skutterudites with the LaFe4Sb12 structure type are one example of a phonon-glass electron-crystal material. Filled skutterudites have a cubic framework of late transition metals, typically Fe or Co, in distorted octahedral coordination by pnictogen atoms, typically Sb, where each pnictogen atom also bonds to adjacent pnictogen atoms.9 A “rattler” atom occupies the large void at the origin of the body-centered unit cell (space group Im
No. 204).9,10 The rigid, covalently bonded framework provides a robust, conductive skeleton needed to satisfy the requirements of an electron crystal, and the large thermal displacement parameters of caged rattler atoms allow them to scatter heat-carrying phonons, decreasing thermal conductivity as in a phonon glass.11–14
Skutterudites with high zT are typically n-type semiconductors with multiple types of filler atoms and a Co–Sb framework.11 The p-type skutterudites also show excellent performance with zT values above 1. High-performance p-type skutterudites typically have a framework of Fe, Co, and Sb with a combination of lanthanide fillers (La–Nd, Gd, Yb).11,12 Using a substantial number of literature examples,11 one finds the best performing p-type skutterudites have a valence electron count (VEC) for one formula unit between 71.5 and 72. Applying this electron counting formalism to LaFe4Sb12, which contains one La3+ ion donating three electrons, four 4s23d6 Fe atoms, and twelve 5p3 Sb atoms results in a VEC of 1(3) + 4(8) + 12(3) = 71. Only electrons at the top of the valence band are accounted for, thus determining the position of the Fermi level, hence, Sb contributes 3 electrons to the count and not 5.11–15 The Sb 5s2 electrons states are part of the valence band but are much lower in energy than the Fermi level.
Using the Zintl–Klemm formalism16,17 and considering that each Sb atom has bonding interactions with two other Sb atoms in the skutterudite structure (Sb–Sb distances ≈3 Å), the following Zintl charges can be assigned: (La+3)(Fe+2)4(Sb−1)12, 1 electron away from charge-balanced composition, consistent with VEC = 71, not 72. For a transition metal (Fe) in an octahedral ligand field of Sb ligands, the low-spin d6 state is the most stable. This makes the replacement of Fe (4s23d6) with Co (4s23d7) or Ni (4s23d8) a viable means of electron-doping to retain the favorable electronic properties of the parent “unfilled” CoSb3 ≡ (Co+3)4(Sb−1)12 skutterudite. Doping Co (or Ni) into Fe sites allows one to adjust the position of the Fermi level and to tune carrier concentrations for optimal electronic transport properties. Although most p-type filled skutterudites with high zT values use Fe and Co frameworks to retain a nearly optimal VEC, some examples do use Fe and Ni. These Ni-containing filled skutterudites have also shown comparable zTs of 1 or more,11–15 but there are far fewer examples of them in literature despite ongoing ethical and geopolitical concerns over the use of cobalt.9,11,18
Design-of-experiments (DoE) tools are commercially available software that use statistical models to map the effects of individual parameters on the outcome of any complex process. With planning, these models are more thorough than the mental picture assembled by a human experimenter who can only track changes made to one variable per experiment.19 By enabling an experimentalist to change more than one variable at a time while exploring multinary phase space,20 statistical design of experiments techniques can vastly decrease the number of attempts needed to find true optima in a given synthetic system.21 By deconvoluting the effects of synthetic variables, one can more fully understand which aspects of a synthetic system are most important to achieving the desired outcome.22–24 Despite the powerful analytical advantages afforded by DoE, its applicability has limits. If an experimenter's intuition can lead to desirable results in fewer trials than DoE requires to model a synthetic system, there is little reason to adopt statistical analysis. This is especially true if the synthesis in question is so expensive or slow that a total of 20 to 30 optimization attempts is impractical.
As such, traditional “heat-n-beat” solid-state synthesis is often unsuitable for statistical DoE. For example, skutterudite synthesis takes about a week by traditional solid-state methods, with two separate heating steps and intermediate regrinding.10–15,25–30 Numerous organic syntheses31–34 and some colloidal nanoparticle syntheses19,21,35 have been optimized using DoE, but all make use of solution-phase methods that can be completed within a few hours. DoE was used to optimize thermoelectric performance in AgSbTe2 by doping, but the general synthetic conditions were unchanged throughout the study.36 Some phase change studies of shape-memory alloys22–24 have been performed using DoE, but the phase changes studied therein occur on a timescale much shorter than a traditional week-long solid-state synthesis. To the best of our knowledge, there are no published examples of solvent-free chemical synthesis processes optimized using statistical DoE tools.
Herein, we applied a fast hydride synthesis route to Ni-doped (Ba/Sr)Fe4Sb12 skutterudites. By making use of commercially available software, for the first time, we use a design of experiments tool to optimize solid-state synthesis, yielding single-phase products in a multi-component system. We further evaluated the thermoelectric properties of the synthesized skutterudites in the context of valence electron counting, carrier concentrations, and thermal transport properties influenced by filler identity and framework doping, achieving a peak zT of 0.54 at 673 K for Sr0.92Fe3.28Ni0.70Sb12.
All weighing of chemicals was performed in an argon-filled glovebox (LC Tech, O2 < 0.5 ppm) and all ball-mill vials were sealed inside two polyethylene bags to retain the inert atmosphere during mixing or crushing. Barium hydride (BaH2, 99.7%, Materion), strontium hydride (SrH2, 99.5%, Materion), iron (Fe, 99.998%, Alfa Aesar), and nickel (Ni, 99.996%, Alfa Aesar) were acquired as powders and used as received. Antimony shot (Sb, 99.9999%, Alfa Aesar) was pulverized to fine powder by two 30 min long rounds of ball-milling in a SPEX 8000M ball mill using a polycarbonate grinding vial with tungsten carbide cap inserts and tungsten carbide balls. Powders of BaH2 or SrH2, Fe, Ni, and Sb were loaded into a polycarbonate ball mill vial with a methacrylate or polycarbonate ball at the bottom for mixing in the ball mill. Mixed powders were brought into a glovebox (LC Tech, <1 ppm O2) equipped with water-cooled TIG welding equipment and transferred into 10 mm diameter by 50 mm long Ta crucibles, which were sealed by arc welding. Once hermetically sealed, crucibles were removed from the glovebox and transferred into a two-part reusable silica reactor. The reactors were evacuated below 30 μbar before loading into the furnace to protect the crucibles inside from oxidation at elevated temperatures. All samples were heated to their target temperatures in 6 hours and annealed for 16 hours. The furnace was switched off, and samples cooled in the furnace with the door closed. Once cool, the crucibles were cut open in the glovebox. Products were usually fine grey powders, although some larger polycrystalline lumps formed sporadically. After grinding gently, samples were stored in the glovebox. Small portions of each sample were removed from the glovebox for characterization.
The duration of ball-milling depended on sample size: 6 minutes for 0.5 gram samples and 18 minutes for the 2.2 gram samples that were used for thermoelectric property measurements. Fe and Ni content was varied around the stoichiometries Sr0.92Fe2.98±xNi1.00±xSb12 or Ba0.9Fe3.00±xNi1.00±xSb12 where x is 0.15 or 0.30. The total sum of Fe and Ni stoichiometry is 4 in Ba0.9Fe3.00±xNi1.00±xSb12, and 3.98 in Sr0.92Fe2.98±xNi1.00±xSb12. For Ni-rich Ba-filled skutterudite samples, less BaH2 was used, their compositions were Ba0.85Fe2.85Ni1.15Sb12 and Ba0.85Fe2.70Ni1.30Sb12. Ba-containing samples with 3 or more equivalents of Fe were heated to 855 K on the 0.5 g and 2.2 g scale. Samples of the composition Ba0.85Fe2.85Ni1.15Sb12 and Ba0.85Fe2.70Ni1.30Sb12 and all Sr-containing samples were heated to 893 K.
Samples of Ba0.9Fe3.00±xNi1.00±xSb12 with 3 or more equivalents of Fe were sintered by heating to 773 K over 10 minutes, followed by 10 minutes of dwell time at 773 K using 90 ± 2 MPa of pressure. Samples of Ba0.9Fe3.00±xNi1.00±xSb12 with less than 3 equivalents of Fe were sintered by heating to 873 K in 11 minutes with a 10 minute dwell under 100 ± 2 MPa of pressure to remove a small impurity of BaSb3, discussed further in the ESI.† Sr-containing samples with 2.68–3.28 equivalents of Fe were sintered by heating to 773 K in 10 minutes and dwelling at 773 K for 10 minutes. 90 ± 2 MPa of pressure were used for the other pellets except Sr0.92Fe2.83Ni1.15Sb12, which was compressed with 100 ± 2 MPa of pressure. Sr0.95Fe3.43Ni0.55Sb12 was sintered by heating to 673 K in 10 minutes, followed by a 10 minute dwell at 673 K and 168 MPa, the maximum pressure possible for a 12.7 mm pellet in the available tooling.
Experimental data produced during the optimization process for BaFe3NiSb12 was input into a response-surface model (RSM) with parallel screening of variables. To use DoE methodology for statistical analysis within JMP, the “custom design” tool was used with uncontrolled variables. The loading stoichiometries of all precursors and dwell temperatures were inherently constrained to experimentally attempted values, defined in Table 1. In addition to these factors, statistical analysis used the quadratic interaction terms Ba stoich. × temp., Ba stoich. × Fe stoich., Ba stoich. × Ni Stoich., Fe stoich. × Ni stoich., Fe stoich. × temp., and Ni stoich. × temp. Antimony loading stoichiometry was set at 12 to normalize the stoichiometry of the other reagents and was not optimized in the statistical model. “Purity” was the phase fraction of skutterudite determined by full-profile fitting using Match! 3 software.37 To ensure accuracy, phase fractions (wt%) were also determined by Rietveld refinement using GSAS-II software38 and found to be consistent (±2–3 wt%) with those determined using Match! 3. Over 100000 random starts were used to optimize the model.
BaaFebNicSb12 | |
---|---|
Factors | Ranges |
Ba stoichiometry, a | 0.71 ≤ Ba ≤ 1.22 |
Fe stoichiometry, b | 2.71 ≤ Fe ≤ 3.03 |
Ni stoichiometry, c | 0.95 ≤ Ni ≤ 1.31 |
Dwell temperature, T (K) | 773K ≤ T ≤ 973 K |
Response: purity | Maximized within the range 0–100% |
The synthetic optimization of SrFe3NiSb12 was performed using the JMP software package to generate a response surface model (RSM) with parallel screening of variables. One center point was used. Constraints were implemented based on preliminary reactions. Table 2 lists the allowed ranges for each factor. In addition to these factors, DoE used the quadratic interaction terms: Sr stoich. × temp., Sr stoich. × Fe stoich., Sr stoich. × Ni Stoich., Fe stoich. × Ni stoich., Fe stoich. × temp., and Ni stoich. × temp. Prior experience dictated that a slight excess of Sb may be needed to account for side-reactions with the crucible walls. Therefore, loading stoichiometries of Fe and Ni were constrained so that total transition metal stoichiometry remained between 3.97 and 4.0 molar equivalents, normalized to 12 molar equivalents of Sb. Relative phase fractions (wt%) were estimated via full-profile fitting using Match! 3 software.37 2224 random starts were used to optimize the model.
SraFebNicSb12 | |
---|---|
Factors | Ranges |
Sr stoichiometry, a | 0.7 ≤ SrH2 ≤ 1.1 |
Fe stoichiometry, b | 2.7 ≤ Fe ≤ 3.3 |
Ni stoichiometry, c | 0.7 ≤ Ni ≤ 1.3 |
Dwell temperature, T (K) | 873K ≤ T ≤ 933 K |
Fe + Ni (constraint) | 3.97 ≤ b + c ≤ 4.0 |
Response: purity | Maximized within the range 0–100% |
![]() | ||
Fig. 1 Overview of the preparation of skutterudites as performed herein. Samples were heated to their annealing temperature within 6 hours and annealed for 16 hours. |
The HT-PXRD data used to study the reaction between BaH2, Fe, Ni, and Sb is shown in Fig. 2. Upon initial heating, all Bragg peaks shift progressively to lower Q values, indicating expansion of the unit cells of unreacted precursors with increasing temperature. BaSb3 and marcasite-type FeSb2 form at 616 K (Fig. S1†). A skutterudite phase begins to form at 755 K together with small amounts of marcasite-type NiSb2 (Fig. S2†). Once the temperature exceeds 835 K, Bragg peaks for all the binaries except NiSb2 disappear. At 855 K, the skutterudite phase is the major crystalline phase with only a trace amount of Fe0.75Ni0.25Sb2 present (Fig. S2†). The skutterudite decomposes into Fe0.75Ni0.25Sb2 and NiSb at 916 K (Fig. S2†). Above 936 K, only the broad, elevated background of the melt and NiSb Bragg peaks match known phases until cooling through 890 K (Fig. S3†). Ni5Sb2 and marcasite-type Fe0.75Ni0.25Sb2 crystallize when cooled below 890 K, with some Sb and the extant NiSb. A skutterudite phase, Ba1.5Sb3, Sb, and marcasite-type Fe0.75Ni0.25Sb2 are observed at 850 K and below as cooling continues (Fig. S3†). No changes in phase or further chemical reactions are visible during cooling below 850 K.
From the in situ data, we can conclude that the filled skutterudite phase BaFe3NiSb12 forms from a mixture of unreacted Sb with BaSb3, FeSb2, and NiSb2 binary intermediates produced during heating. This reaction pathway offers an important hint as to why the exact Fe/Ni ratio is not very important to the purity of the final skutterudite phase, enabling the synthesis of BaFe3.30Ni0.70Sb12 at the same temperature as BaFe3NiSb12. Based on in situ diffraction data, FeSb2 and NiSb2 act as the primary Fe and Ni sources for skutterudite formation. Phase analysis of ex situ powder X-ray diffraction data from synthesis attempts at various temperatures and Ba stoichiometries offers further insight into the reaction mechanism and reproducibility of 2.2 g-scale reactions when making Ba0.9Fe3.00±xNi1.00±xSb12.
Reactions in the laboratory occur under different conditions than in situ reactions. However, the same intermediates and decomposition products are shown in the ex situ PXRD patterns, such as the BaSb3 resulting from a small excess of BaH2 in Fig. 3A. Ex situ reactions above the optimal temperature produce NiSb and Fe0.75Ni0.25Sb2 impurities due to decomposition of the skutterudite, as in Fig. 3B. NiSb is seen throughout the HT-PXRD data as a decomposition product. In ex situ diffractograms, NiSb is also seen in samples that were not heated to a high enough temperature. NiSb is quite thermally stable, so it is likely the last binary to react and the first decomposition product to form. Therefore, the presence of NiSb alone in a sample is not diagnostic of synthesis temperatures being too low or too high. When NiSb is accompanied by BaSb3 and binary marcasites FeSb2 and NiSb2, the combination of impurities indicate that synthesis temperatures were too low, as these binaries would further react to form a skutterudite phase at higher temperatures. Samples with NiSb and Fe0.75Ni0.25Sb2 together indicate that the annealing temperature was too high, leading to skutterudite decomposition.
Increased Ni content in the samples with loading compositions Ba0.90Fe2.70Ni1.30Sb12 and Ba0.90Fe2.85Ni1.15Sb12 resulted in an increased phase fraction of NiSb, necessitating the increase in optimal annealing temperature from 855 K (Fig. 3A) to 893 K (Fig. 3C) in 2 gram reactions. Excessive BaH2 content results in a BaSb3 impurity, which has visible Braggs' peaks in the Ba-rich samples (Fig. 4A, C and G). Temperatures below 855 K result in binary and transition metal-antimony compounds, most of which are marcasite-type FeSb2 and NiSb2. The reaction is temperature-dependent enough that samples annealed within less than 50 K are distinguishable by PXRD. Samples of identical composition annealed at 773 K (Fig. 4F and L) and 823 K (Fig. 4E and K) have different impurities than samples annealed at 873 K (Fig. 4D and J).
This is most likely why examples of DoE tools applied to inorganic solid-state materials are rare and tend to involve somewhat fast solution-phase syntheses,21,35 phase transitions,24 properties optimization,36 or device fabrication.52 After all, performing approximately 20–30 experiments to reach one definitive answer makes DoE unappealing when traditional solid-state reactions often take a week or longer.25–30 The sluggishness of many solid-state reactions is caused by the slow diffusion of solids. Fortunately, solid-state syntheses using hydride precursors are highly reproducible and reasonably fast, with typical durations of approximately 24 hours and no need to regrind and reheat samples.42 Therefore, quick reactions enabled by hydride precursors make the requisite number of reactions for statistical DoE much less daunting.
A pace of two days to complete three reactions simultaneously makes it possible to gather enough data in a short time to use DoE when performing solid-state synthesis using hydride precursors. To validate the utility of DoE for solid state synthesis, we used previously collected synthesis data and modeled it using uncontrolled continuous variables within the ranges listed in Table 1. Because this is an analysis of existing data and not new experiments, this constitutes a statistical analysis rather than a DoE. Had the data been collected at pre-planned, statistically optimal data points, this would be DoE. A response surface model (RSM) is an nth dimensional surface that relates the values of independent variables, called factors, to the value(s) of the dependent variables, called responses, where n is the “level” (quadratic functions are 2-level, cubic functions are 3-level, etc…) of the model. A 2-level response surface model based on Mayer and Nachtsheim's coordinate-exchange algorithm40 was used herein. The coordinate-exchange algorithm uses iterative “random starts” to optimize a model. To do this, randomly chosen values of the constants associated with each term in the equation that defines the response surface are replaced with values at the extrema of the surface, and the version of the equation with the lower variance is kept. The global optimum is the equation with the lowest variance. In non-trivial cases, at least 1000 random starts are needed, with greater numbers of starts increasing the chances of finding the globally optimal model.40,41 There are other model types that are more predictive than RSMs or that require fewer data points, but they would require more data or generate less predictive results, respectively. The software used here, JMP Pro 16 (Statistical Discovery LLC), automatically performs screening of variables (analyzes the importance of all variables to each outcome), highlighting which factors are most important to achieving a desired result. This screening is especially useful for subsequent rounds of optimization19 because further iterations of a model can be made without unimportant factors, and models with fewer variables require fewer experiments.
Sampling as many different regions of the modeled synthetic space as possible was especially important to the validation of JMP using data shown in Fig. 4 and S4.† These data were collected during attempts to optimize the bulk synthesis of BaFe3NiSb12 before in situ data was collected. The model based on these data (Fig. 5) covers the range of temperatures and compositions listed in Table 1 and shown in the loading ratios and annealing temperatures illustrated in Fig. 4 and S4.† The in situ PXRD data herein reveal the reaction onset temperature and demonstrate that reactions proceed to near completion within seconds. Based on our prior experience with hydride reactions of pnictides, reactions achieve thermal equilibrium within a few hours of reaching a given temperature, and a fixed heating rate can be used.42–48 As such, a 16-hour annealing time was deemed to be sufficient to ensure complete reaction and was kept constant across all synthesis attempts.
Limited to the results from previous experimental attempts, every available piece of data had to be used. The PXRD patterns of each product are shown in Fig. 4 and S4.† Fig. 5 and Table 3 show the results of this statistical analysis model. The “desirability” of a factor represents what values that factor can take to achieve an optimal result. If all factors have a desirability of 1, the optimal result should occur.53 The “LogWorth” is the negative log10 of the p-value.54 As the p-values decrease, LogWorth increases, and the more important that factor is to the outcome of an experiment54 (see ref. 53 for an in-depth discussion of the p-value). If LogWorth is above 2, as indicated by the blue vertical lines in Table 3, changes to the factor in question have a statistically significant impact on the response. If LogWorth is under 2, the factor might not be changing the response. If no factor has a LogWorth above 2, something needs troubleshooting: perhaps the model is ill-defined, the dataset is biased or too small, or the process undergoing optimization is irreproducible. In our case, the dataset is biased. The data significantly oversamples stoichiometry of approximately Ba0.8Fe3.0Ni1.0Sb12 with annealing at 873 K. Grey surfaces in Fig. 5 are the confidence intervals for each factor. The further the grey area extends from the black line inside it, the larger the confidence interval, and the less precise (and potentially less accurate) the model is. The pinched shapes of grey surfaces (Fig. 5) in the purity vs. factor plots and the high importance of the Fe stoich. × Ni stoich. quadratic term (Table 3) primarily indicate the biases in the data. Fe and Ni were exchanged for one another during synthesis because the target was a solid solution of skutterudites Ba0.9Fe3.00±xNi1.00±xSb12, therefore, the sum of Fe and Ni stoichiometry must be 4, the required stoichiometry to form skutterudites.
The importance of Fe and Ni stoichiometry was exaggerated relative to temperature and Ba stoichiometry because of the exchange of Fe and Ni to maintain a total stoichiometry of 4. The ratio between Fe and Ni is known not to matter because any Fe/Ni can be used within the defined limits when performing reactions on a half-gram scale. Fe–Ni stoichiometric variations had to be modeled because the only samples annealed under 873 K had varying Fe and Ni stoichiometries. Therefore, the possibility of Fe and Ni stoichiometry having a meaningful effect on synthesis temperature could not be ignored. On the half-gram scale, no difference in optimal annealing temperature was noted across the Ba0.9Fe3.00±xNi1.00±xSb12 compositions. All data used to make the model was taken from 0.5 g scale reactions. Therefore, the model had no way to account for the need to use less BaH2 and anneal Ni-rich Ba0.9Fe3.00±xNi1.00±xSb12 samples at a higher temperature than Fe-rich samples in reactions performed on the 2.2 gram scale (see Experimental for synthesis details). The synthetic conditions recommended by the model were a sample with loading composition 0.95 BaH2 + 2.92 Fe + 1.03 Ni + 12 Sb annealed at 861 K. This is remarkably close to the optimal compositions used to make samples, especially the most important – temperature. The accuracy of this model in predicting successful synthesis conditions highlights the importance of preliminary reactions and well-chosen ranges for each parameter when setting the limits of a model. The center of the modeled space (average of the high and low limits) should be near the expected optimal value so that the software can calculate meaningful regression statistics on all sides of the global minimum in the difference between the regression (model) and experimental data. Furthermore, no result can be recommended outside the space in which the model is defined. Therefore, one must encompass the optimal reaction conditions within the modeled space without making the range of possible conditions so large that meaningful results can hide in the empty space between data points.
The loading stoichiometry (Ba0.90Fe3.00Ni1.00Sb12) was chosen based on the VEC formalism and trial-and-error. Annealing at 855 K was chosen based on HT-PXRD experiments. Ba0.90Fe3.00Ni1.00Sb12 samples annealed at 855 K result in samples composed of the desired skutterudite phase and a miniscule amount of unreacted Sb. Given a small and imperfect dataset, the model's ability to come within a margin of 6 K from an experimentally measured optimal temperature is very impressive (861 K vs. 855 K) within the total range of 200 K. Therefore, the authors contend that the model is successful despite sampling biases in the experimental data points selected for the model (Fig. 4 and 5). These biases (and lack of constraints) caused the deviation from optimal stoichiometry. In short, the data set chosen for this model oversampled annealing at 873 K. There were no samples annealed below 873 K with Fe stoichiometry significantly different from 2.7 and Ni stoichiometry differing from 0.7. As a result, the model cannot fully deconvolute the importance of temperature from the importance of Ni stoichiometry. However, this sampling bias would not hinder carefully designed optimization processes that use DoE from the beginning.
By analogy to Ba0.9Fe3.00±xNi1.00±xSb12, we postulate that exact Fe and Ni stoichiometry should not matter if the sum of Fe and Ni stoichiometries remains between 4 and 3.97. However, Fe and Ni stoichiometries were modeled in case Sr0.92Fe2.98±xNi1.00±xSb12 was less analogous to Ba0.9Fe3.00±xNi1.00±xSb12. The sum of Fe and Ni stoichiometries was confined to the range 3.97 ≤ total ≤ 4.0 in the case of DoE optimization of Sr0.92Fe2.98±xNi1.00±xSb12 synthesis. Sr stoichiometry was modeled in the range from 0.7 to 1.1 equivalents, with the lower boundary corresponding to the minimal probable filling fraction of Sr into the central voids of the skutterudite structure as estimated by SEM/EDXS, while the upper boundary was selected based on the maximum equivalents of SrH2 that do not result in excessive phase fractions SrSb2 impurity detectable by laboratory PXRD. Modeling the strontium stoichiometry is important because insufficient filler metal will result in deviations from the nominal valence electron count and decrease the number of rattler atoms available to dissipate phonons. Excess Sr causes the formation of Sr–Sb air-sensitive binary impurities that could alter the properties of a pellet. The “boxes” defined by factors (Tables 2 and 4) that encompass the phase space within the RSM model are illustrated in Fig. 6. The model was designed with a center point (experiments at the exact center of the modeled “box”) to examine the reproducibility of the process, and 15 screening points (experiments at varying conditions) were used to find an experimental “best guess” as to what optimal synthesis conditions should be. We did not use replicate trials because we decided that conserving materials and getting results quickly was more important than assuring the statistical significance of our results, especially considering the accuracy of the model of the Ba-filled skutterudites considered earlier.
The suggested optimal conditions (red lines in Fig. 6A and B) indicate that 0.81 to 1.05 equivalents of SrH2 should be optimal, depending on Ni stoichiometry, and an annealing temperature of 893 K is preferred at all compositions. Only 2224 random starts of the coordinate exchange algorithm were used to optimize four continuous variables with one response, which is rather few,40,41 and the results were statistically insignificant. Nevertheless, the generated model remains very useful. The most important terms were Sr stoichiometry and the quadratic term relating Sr stoichiometry and annealing temperature (Table 4). The next most important terms are the Sr stoichiometry itself and the quadratic terms that involve Sr stoichiometry. This is expected based on the analogy to the statistical analysis of Ba0.9Fe3.00±xNi1.00±xSb12 synthesis.
Technically, the results of DoE optimization were statistically insignificant, as LogWorth was below 2 for all parameters (left of the blue line in Table 4). Hence, optima were predicted to exist at the extrema of Fe and Ni stoichiometry. However, careful analysis of the model and PXRD data (Fig. S5 and S6†) allows one to extract useful information from this round of DoE. Sr0.92Fe2.98±xNi1.00±xSb12 is part of a solid solution where x is small, so Vegard's rules should apply. Following Vegard's rules, if 893 K is the optimal annealing temperature for Fe-poor (Fe2.67Ni1.30) and Fe-rich (Fe3.32Ni0.678) compositions, 893 K should also be optimal for the stoichiometries between the extrema. By examining the secondary phases in our reaction products, we can validate the predicted optimal temperature. Ternary marcasite Fe1−xNixSb2 forms when annealing at 903 K and 933 K, and binary marcasites NiSb2/FeSb2 form at 873 K (Fig. S5 and S6†). Binary marcasites are precursors to skutterudite, while ternary marcasite is a decomposition product. Therefore, the optimal temperature for skutterudite synthesis must be between 873 K and 903 K, so the temperature of 893 K predicted by DoE is a reasonable result. Therefore, with the right amount of SrH2, the annealing temperature should be set to 893 K for any Fe/Ni ratio. The DoE model suggests that the Fe-rich (Fe3.32Ni0.678) composition requires 1.05 equivalents of SrH2, and the Fe-poor composition (Fe2.67Ni1.30) needs only 0.81 equivalents of SrH2 (Fig. 6). Following Vegard's rules, a sample with the composition Sr0.92Fe2.98Ni1.00Sb12 annealed at 893 K should be pure. Exactly such a sample showed no binary or ternary impurities on the 2 gram scale (Fig. S7†). Ultimately, regardless of Fe and Ni stoichiometry, 0.92 equivalents of SrH2 provided pure Sr0.92Fe2.98±xNi1.00±xSb12 samples with x ≤ 0.3. On the 0.5 g scale, laboratory PXRD experiments showed the products contained minor impurities of Sb, FeSb2, and NiSb2 in addition to skutterudite (Fig. S7†). Further reactions revealed that this scale dependence is merely a case of transfer losses affecting the least plentiful reagent disproportionately, discussed further in the ESI (Fig. S7†). Because 2 gram samples are required to press pellets for high-temperature thermoelectric property measurements, synthesis at the 0.5 gram scale was not pursued further.
Based on the authors' experiences with statistical DoE, we recommend the following steps when approaching an unfamiliar synthesis using statistical DoE as a guide. After a literature survey, the next step is a series of preliminary reactions guided by chemical intuition, just enough to find conditions in which the target compound can form. Next, one uses the knowledge acquired through the literature survey and preliminary reactions to choose factors for the DoE model. Permissible ranges for each factor are based on the results of preliminary reactions. The more factors there are in a model, the more trials are needed. By creating mock RSM DoE models in JMP software with various designs, we selected the one requiring the fewest data points, resulting in 16 reactions total. To ensure the resulting model is representative of the system in question, the reaction conditions must be attempted with the exact parameters suggested by the statistical software. After the model predicts optimal conditions, a batch of reactions at and near the predicted conditions should be attempted. If the predicted optimum does not produce the desired result, one must perform another iteration or choose different parameters to optimize. This will most likely entail removing unimportant factors from the model, narrowing the range of values for some factor(s), or both.
Sometimes, the unimportance of certain aspects of synthesis can be exploited. In our case, the less important variables in the model of Ba0.9Fe3.00±xNi1.00±xSb12 and Sr0.92Fe2.98±xNi1.00±xSb12 were the exact loading stoichiometries of Fe and Ni. The ability to substitute Fe with Ni enabled us to tune the carrier concentrations (Fig. 7 and S8†) and improve thermoelectric performance.
Crystallographic densities were calculated using the refined unit cell parameters (Fig. 7) of Ba0.90Fe3.00Ni1.00Sb12 for the Ba-series and Sr0.92Fe2.98Ni1.00Sb12 for the Sr-series. The compactness of pellets above 93% (Table S1 and Note S1†) of crystallographic density was achieved using individually tailored sintering conditions. The skutterudite phases were uniform in composition based on EDXS, although some SrO and BaO grains were seen at pellet surfaces (Fig. S9†) which resulted in scratches during polishing (Fig. S10†). Sintering results for pellets of all compositions (Table S1†) and backscattered electron micrographs (Fig. S9†) indicate that the pellets have similar microstructures and are compact enough for their transport properties to be directly comparable. Skutterudite grains in the SPS pellets are about 5 μm across and gain no obvious crystalline habit. Grain boundaries show little porosity and do not appear to contain precipitates.
EDXS measurements of transition metal stoichiometries are within one standard deviation of the stoichiometry used to make the sample in question. Skutterudites almost always have fractional occupancy of filler atoms.11,12,27,28 EDXS suggests that the Sr-series have filling fractions of 80% (Table S1†), and the Ba series have filling fractions in the range of 76–90% that decrease when more Ni is added (Table S1†). The authors hypothesize that the Ba-series compensates for electron deficiency in the framework by accepting more electron-donating filler atoms into the structure. This variance in filling fraction with aliovalent doping of the skutterudite framework is also supported by the variation in SrH2 and BaH2 used to make pure skutterudites. Ni-enriched (hole deficient) Ba0.85Fe2.85Ni1.15Sb12 and Ba0.85Fe2.70Ni1.30Sb12 need only 0.85 instead of 0.90 equivalents of BaH2 to minimize the phase fraction of BaSb3 impurity in the reaction product. Ni-deficient (electron deficient) Sr0.95Fe3.43Ni0.55Sb12 is the only Sr-filled skutterudite herein that needs more than 0.92 equivalents of SrH2 to minimize marcasite-type impurities (Note S2†).
Because of their similar VECs (Table S1†) and concomitant metallicity, the positive (p-type) Seebeck coefficients, S(T), of Ba0.90Fe3.30Ni0.70Sb12, Ba0.90Fe3.15Ni0.85Sb12, and Ba0.90Fe3.00Ni1.00Sb12 reach maximum values between +135 and 150 μV K−1. As shown in Fig. 8A and C, the semiconductor-like skutterudite in the series, Ba0.85Fe2.85Ni1.15Sb12, has a lower peak Seebeck coefficient of 87 μV K−1. With a valence electron count of 72.3 (Table S1†), Ba0.85Fe2.70Ni1.30Sb12 is over-doped with Ni and displays n-type charge transport, indicated by its negative Seebeck coefficient. The magnitudes of the peak Seebeck coefficients observed for the Ba-filled skutterudites are limited by the excitation of minority charge carriers. As temperature increases, electrons are excited into the conduction band in the p-type samples, as were holes in the n-type compound, resulting in bipolar charge transport. This is not unusual for thermoelectric materials measured near the top of their useful temperature range.
The skutterudites reported herein demonstrate bipolar conduction at temperatures lower than skutterudites made with Fe and Co that have the same VEC.11 The extent of bipolar charge transport observed herein is consistent with the previous report of Ba0.96Fe3.0Ni1.0Sb12.28 The values of the Seebeck coefficient at various temperatures are not directly comparable between that study28 and this report. Here we used a Netzch Nemesis SBA, which measures the Seebeck coefficient without inducing a large thermal gradient across the sample. However, the previous report of Ba0.96Fe3.0Ni1.0Sb12 reports a Seebeck coefficient measured with an ULVAC-Rico ZEM-3 instrument, which has a different sample environment, resulting in a large temperature gradient across the sample and causing a pronounced cold-finger effect and significant overestimation of the Seebeck coefficient.55 The temperature of the inflection points in the curves are comparable however, and quite similar. The highest peak Seebeck coefficient in the Ba-series herein belongs to a bad metal, not a heavily doped semiconductor. Therefore, the most important result of aliovalent doping in Ba0.9Fe3.00±xNi1.00±xSb12 is the suppression of bipolar charge transport. This is demonstrated by the two Ba0.9Fe3.00±xNi1.00±xSb12 with the best electronic transport performance. Ba0.90Fe3.15Ni0.85Sb12 had the highest peak Seebeck coefficient of the Ba-series with 146 μV K−1 at 575 K despite its low VEC of 71.3. The onset temperature of bipolar charge transport is further increased for Ba0.90Fe3.30Ni0.70Sb12; its Seebeck coefficient plateaus at 675 K instead of 575 K like Ba0.90Fe3.15Ni0.85Sb12. Overall, increasing Ni content in our Fe–Ni skutterudites decreases the peak value of the Seebeck coefficient. Electrons from Ni cause excessive bipolar charge transport at lower VECs than would be expected for skutterudites made with a Fe–Co framework, as seen in literature.11,49 Our results indicate that the optimal range of VECs for p-type Fe–Ni skutterudites is 71.0–71.5 rather than 71.5–72.0 as in Fe–Co skutterudites.
Thermal conductivities, κ(T), are all similar for Ba-filled skutterudites (Fig. 8B), approximately 3 W mK−1, consistent with the values previously reported49 for Ba0.96Fe3.0Ni1.0Sb12. All values of thermal conductivity in the Ba-series are within experimental error except Ba0.90Fe3.30Ni0.70Sb12, which is more conductive near room temperature. All thermal conductivity curves also have similar shapes. One of the most noticeable differences between each curve is the position of the inflection point. In the p-type skutterudites pictured above, the temperature of the inflection points in κ(T) increase as Fe content increases, which is attributed to the suppression of thermal conduction due to bipolar charge transport. The electronic contributions to thermal conductivities were calculated based on the Seebeck coefficients using Lorenz numbers calculated with the empirical equation proposed by Snyder et al.56 The differences between total thermal conductivities above are also caused by electronic contributions to thermal conductivity, shown in Fig. S12A.† On average, the lattice thermal conductivity of the Ba-filled skutterudites discussed here is approximately 2.5 W mK−1, which is common for singly filled skutterudites.11 The lattice contribution to thermal conductivity is greater than the electronic component in all cases and accounts for roughly 3/4 of the total thermal conductivity. Electronic contributions to the total thermal conductivity of p-type samples show no trend with respect to Fe or Ni content or the valence electron count (Fig. S12A†).
Among our p-type Ba0.9Fe3.00±xNi1.00±xSb12 skutterudites, zT increased with Fe content. The highest zT of 0.34 is achieved by Ba0.90Fe3.30Ni0.70Sb12 at 675 K, and the temperature-dependence of zT did not plateau or decrease in the measured temperature range. The curves in Fig. 8D never cross over one another, so there is no temperature within the range tested in which Ba0.90Fe3.30Ni0.70Sb12 is not the highest-performing Ba-filled skutterudite in this study because it was the least encumbered by bipolar charge transport. However, further doping of Fe in place of Ni would not be productive for Ba0.9Fe3.00±xNi1.00±xSb12 skutterudites, as indicated by the similarities of the Seebeck coefficients of Ba0.90Fe3.15Ni0.85Sb12 and Ba0.90Fe3.30Ni0.70Sb12. More holes will not result in better thermoelectric performance, as shown by the carrier concentration of Ba0.90Fe3.30Ni0.70Sb12 (Fig. S8†), which exceeds 1.0 × 1020 cm−3, the desirable carrier concentration in skutterudites.11,15
Comparison of Sr0.92Fe2.98±xNi1.00±xSb12 and Ba0.9Fe3.00±xNi1.00±xSb12 reveals intriguing trends with regards to thermal conductivity. Like the Ba-series, p-type Sr0.92Fe2.98±xNi1.00±xSb12 skutterudites show inflection points in their κ(T) curves (Fig. 9B). These inflection points also shift to higher temperatures as compositions get more Fe-rich and the bipolar contribution to thermal conductivity is minimized. Thermal conductivities for Sr0.92Fe2.98Ni1.00Sb12 and Sr0.92Fe2.83Ni1.15Sb12 are significantly lower than for the other skutterudites herein (≈2.5 W mK−1 at room temperature rather than ≈3.0 W mK−1). Based on the lack of porosity in the sintered pellets observed by SEM (Fig. S11†) and the similarity between Archimedean and geometric densities of the pellets (Table S1†), low compactness of the sintered pellets is not the reason for their low thermal conductivity. Lattice and electronic contributions to total thermal conductivity (Fig. S12B†) just happen to be in a “sweet spot” for Sr0.92Fe2.98Ni1.00Sb12 and Sr0.92Fe2.83Ni1.15Sb12.
All the p-type skutterudites in the Sr-series have lattice thermal conductivities within experimental deviation of one another (Fig. S12†). Overall, the lattice contributions to thermal conductivity are smaller in the Sr-series (≈2.4 W mK−1) than in the Ba-series (≈2.8 W mK−1). The electronic contributions to thermal conductivity decrease with Ni content across the Sr-series as one would expect when electron-doping a p-type bad metal (Fig. S12A†). The electronic and lattice contributions to the thermal conductivity of Sr0.92Fe3.28Ni0.70Sb12 become very similar (≈1.3 W mK−1 and ≈1.6 W mK−1 respectively) at 675 K, which is unusual for skutterudites, which typically have larger lattice contributions to the total thermal conductivity.
Like the Ba-series, peak thermoelectric figures of merit (zT) for the Sr-filled iron–nickel skutterudites increase with decreasing Ni content until Sr0.92Fe3.28Ni0.70Sb12, which has a peak zT of 0.54 at 673 K. We note that using values of thermal conductivity determined experimentally by reference to a pyroceram standard pellet results in a lower thermal conductivity, and a peak zT of 0.66 at 673 K (κexperimental ≈ 2.3 W mK−1). However, the temperature dependence of the experimentally determined thermal conductivity (Fig. S13†) fluctuates considerably due to variations in the experimentally determined heat capacity of the sample.
Sr0.92Fe3.28Ni0.70Sb12 has the highest zT of the materials reported herein, with the next highest being Sr0.92Fe3.12Ni0.85Sb12 and Sr0.92Fe2.98NiSb12. Further decreasing Ni stoichiometry does not increase zT. The differences in values and curve shapes between Sr0.92Fe3.28Ni0.70Sb12 and Sr0.92Fe3.12Ni0.85Sb12 in Fig. 9 were minimal except Fig. 9D. Lower Ni content results in excessive hole concentrations, and therefore, decreased thermoelectric performance. The carrier concentration of Sr0.92Fe3.28Ni0.70Sb12 already exceeds 1.0 × 1020 per cm3 (Fig. S8†), the preferred11 carrier concentration in thermoelectric skutterudites.
Sr0.95Fe3.43Ni0.55Sb12 (made as a proof-of-concept) has a smaller Seebeck coefficient than Sr0.92Fe3.28Ni0.70Sb12 (Fig. S14†). The increased resistivity of Sr0.95Fe3.43Ni0.55Sb12 as compared with Sr0.92Fe3.28Ni0.70Sb12 is likely a result of poor compaction caused by the poor stability of Sr0.95Fe3.43Ni0.55Sb12 during sintering. However, even if Sr0.95Fe3.43Ni0.55Sb12 had the same resistivity as Sr0.92Fe3.28Ni0.70Sb12, its lower Seebeck coefficient would result in a lower zT. The previous report of SrFe4Sb12 reinforces this point. The resistivity and Seebeck coefficient of SrFe4Sb1257 were smaller than Sr0.92Fe3.28Ni0.70Sb12. Other literature also shows that doping of Fe–Ni skutterudites results in maximum zT values across a series of compositions when the VEC is between 71 and 72,13–15,25,26,58 not 70 as in SrFe4Sb12. Based on thermoelectric properties observed in this study and prior literature, future p-type Fe–Ni skutterudites should target a VEC in the range 71.0–71.5 and prefer Sr-filling over Ba-filling.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta03288a |
‡ Present Address: Department of Chemistry, University of Utah, Salt Lake City, Utah 84112, USA. |
This journal is © The Royal Society of Chemistry 2025 |