5,571
Views
32
CrossRef citations to date
0
Altmetric
Focus on Intermetallic Catalysts

The emergence of Heusler alloy catalysts

ORCID Icon, ORCID Icon & ORCID Icon
Pages 445-455 | Received 01 Feb 2019, Accepted 19 Mar 2019, Published online: 10 May 2019

ABSTRACT

For over a century, Heusler alloys (X2YZ; X, Y: transition metals, Z: typical metals) have attracted attention as magnetic materials. Nowadays, they are also popular as thermoelectric and shape memory materials. However, until very recently, Heusler alloys had been almost unknown as catalysts. We conceived that they could be ideal candidates for catalysts because (1) they have so many elemental sets and (2) they can be partially substituted with other elements (e.g. X2YZ1–xZ’x) in many cases. We first investigated a variety of Heusler alloys as catalysts and then discovered effective catalysts without precious metals for the selective hydrogenation of alkynes based on the aforementioned feature (1) and demonstrated a systematic tuning of catalytic properties by applying the feature (2). Heusler alloy catalysts have also recently been studied by other groups; hence, this traditional alloy group is becoming a new stream in terms of catalysis. In this article, we review the current progress on Heusler alloy catalysts and describe their future prospects.

Graphical Abstract

This article is part of the following collections:
Intermetallic CatalystsCatalystsSpintronics / Magnetic Materials

1. Introduction

The words ‘intermetallic compounds’ and ‘ordered alloys’ are often used synonymously. In a narrow sense, intermetallic compounds have no order–disorder transition temperature below a melting point and have a limited allowable composition range, while ordered alloys have the transition temperature and a relatively wide composition range, as shown in [Citation1,Citation2]. In an A-B alloy, an intermetallic compound with highly covalent bonds will form if the A-B bond is stronger than the A-A and B-B bonds, but an ordered alloy with a weakly covalent bonding will form if the A-B bond is moderately stronger. Thus, there are many intermetallic compounds in alloys consisting of transition metals and typical metals. However, many ordered alloys consist of only transition metals. Heusler alloys (Heusler compounds) (X2YZ) typically consist of transition metals in groups 8–12 for X, 3–8 for Y, and typical metals in group 13–15 for Z, as shown in [Citation3Citation5]. Depending on elemental sets of X, Y, and Z, they are typical intermetallics, typical ordered alloys showing transitions of L21 → B2 → A2 (body-centered cubic, bcc; ), or in-between alloys only showing the L21 → B2 transition. Note that there are also D03 phases with disordering between X and Y atoms, half-Heusler, inverse Heusler, and quaternary Heusler alloys, as shown in .

Figure 1. Phase diagrams of (a) Al-Pt system [Citation1] forming intermetallic compounds and (b) Fe-Pt system [Citation2] forming ordered alloys. The diagrams were traced from National Institute for Materials Science (NIMS) AtomWork <http://crystdb.nims.go.jp/> [Citation1,Citation2].

Figure 1. Phase diagrams of (a) Al-Pt system [Citation1] forming intermetallic compounds and (b) Fe-Pt system [Citation2] forming ordered alloys. The diagrams were traced from National Institute for Materials Science (NIMS) AtomWork <http://crystdb.nims.go.jp/> [Citation1,Citation2].

Figure 2. Schematic illustration of the elemental substitution of Heusler alloys (X2YZ) [Citation3] and typical components of X, Y, and Z (highlighted in periodic table). Other elements such as lanthanoids and those in groups 1 and 2 are also possible components (see Ref [Citation4]). The figure was excerpted from Ref [Citation5], ©The Authors, some rights reserved; exclusive license American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC http://creativecommons.org/licenses/by-nc/4.0).

Figure 2. Schematic illustration of the elemental substitution of Heusler alloys (X2YZ) [Citation3] and typical components of X, Y, and Z (highlighted in periodic table). Other elements such as lanthanoids and those in groups 1 and 2 are also possible components (see Ref [Citation4]). The figure was excerpted from Ref [Citation5], ©The Authors, some rights reserved; exclusive license American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC http://creativecommons.org/licenses/by-nc/4.0).

Figure 3. Crystal structures for (a) L21 (full-Heusler, Fm–3m), (b) B2 (Pm–3m), (c) A2 (bcc, Im–3m), (d) D03 (Fm–3m), (e) C1b (half-Heusler, F–43m), (f) Li2AgSb or Hg2CuTi (inverse Heusler, F–43m), and (g) LiMgPdSn (quaternary Heusler, F–43m) phases, where symbols in parentheses represent space groups [3]. In (b), the Y and Z atoms are mixed. In (c), all atoms are mixed. In (d), X and Y elements are mixed. In (e), half of X atoms are absent, so that the structure is called a half-Heusler (XYZ) in contrast to a full-Heusler (X2YZ), in a narrow sense. In (f), additional Y atoms occupy the vacant sites of the half-Heusler. In (g), the fourth elements (A) occupy the vacant sites of the half-Heusler. (f) and (g) show the prototype names because their Strukturbericht symbols (like L21) are not commonly used. For (f), there are two prototype names for the same structure.

Figure 3. Crystal structures for (a) L21 (full-Heusler, Fm–3m), (b) B2 (Pm–3m), (c) A2 (bcc, Im–3m), (d) D03 (Fm–3m), (e) C1b (half-Heusler, F–43m), (f) Li2AgSb or Hg2CuTi (inverse Heusler, F–43m), and (g) LiMgPdSn (quaternary Heusler, F–43m) phases, where symbols in parentheses represent space groups [3]. In (b), the Y and Z atoms are mixed. In (c), all atoms are mixed. In (d), X and Y elements are mixed. In (e), half of X atoms are absent, so that the structure is called a half-Heusler (XYZ) in contrast to a full-Heusler (X2YZ), in a narrow sense. In (f), additional Y atoms occupy the vacant sites of the half-Heusler. In (g), the fourth elements (A) occupy the vacant sites of the half-Heusler. (f) and (g) show the prototype names because their Strukturbericht symbols (like L21) are not commonly used. For (f), there are two prototype names for the same structure.

Heusler alloys have great features. There are so many possible sets of X, Y, and Z (). This feature predicts a discovery of new catalysts, and enables them to substitute each other (e.g. X2YZ1–xZ’x). The substitution of Heusler alloys is common practice in other fields, especially because their electronic structures can be controlled in a rigid band-like fashion [Citation6,Citation7]. Thus, the substitution would be useful for tuning catalytic properties through controlling electronic structures (ligand effect) and also through changing surface elements (ensemble effect), as shown in . However, Heusler alloys were almost unknown as catalysts even though they have been popular as magnetic (spintronic), thermoelectric, and shape memory materials [Citation6Citation8]. Especially, they have attracted attention of magnetic researchers since the discovery of Cu2MnAl over a century ago [Citation9,Citation10] because they display ferromagnetism without ferromagnetic elements. Focusing on these features, we started a study on Heusler alloys as new catalysts, and experimentally investigated their catalytic properties [Citation11,Citation12]. In this study, we roughly revealed the catalytic properties of Heusler alloys by investigating 12 alloys for the hydrogenation of propyne and the oxidation of carbon monoxide. Recently, we discovered effective catalysts without noble metals for the selective hydrogenation of alkynes and achieved a systematic control of catalytic properties by elemental substitution, by using the features of Heusler alloys [Citation5]. In this article, we review our studies in Sections 2 and 3, and briefly refer to other studies on Heusler catalysts in Section 4 [Citation13Citation15]. Finally, we summarize our studies and describe future prospects in Section 5.

2. Experimental details

Details are described in the previous papers [Citation5,Citation11,Citation12]. Alloy ingots were prepared by arc-melting from pure metallic pieces (purity ≥ 99.9%). They were annealed in the Ar atmosphere for homogenizing compositions (typically 1000 °C) and ordering atomic positions (typically 500–600 °C) and then crushed using a mortar and pestle. The powders obtained were sieved to a particle size of 20–63 μm, and then used as catalysts. The sieving was necessary to exclude as many uncertain factors as possible. X-ray diffraction (XRD) was performed for a powder sieved to <20 μm after annealing at 600 °C for 1h to remove the defects introduced by crushing. The XRD confirmed that all samples were (almost) a single phase of the L21 structure with high order. The Brunauer–Emmett–Teller method for Kr adsorption isotherms at 77 K yielded the surface area of catalysts as 0.05–0.13 m2 g–1 mainly depending on atomic weights.

Catalytic tests were conducted using a standard fixed-bed flow reactor built using Swagelok® parts made of 316 stainless steel. An appropriate amount of catalyst (typically 0.4 g) was placed into a quartz tube (4-mm internal diameter) and then inserted into an electric furnace. Before the reaction, gas lines, including gas regulators, were purged enough to remove air contamination accumulated due to a slight leakage at the joints, and the catalyst was heated at 600 °C for 1h under hydrogen flow for removing surface oxides and defects. After the introduction of a reaction gas at room temperature at a flow rate of 30 cm3 min–1, heating was started to measure temperature dependence. Unreacted and produced gases were analyzed using a gas chromatograph.

3. Results and discussion

3.1. Oxidation of carbon monoxide [Citation11,Citation12]

The 12 Heusler alloys listed in were investigated for the oxidation of carbon monoxide using a reactant of [1.2%CO/0.4%O2/He balance], the concentration of which was CO rich to suppress an irreversible oxidation of catalysts. shows conversions of CO during continuous heating by 2.5 °C min–1. In pure X, the hierarchy, Cu < Co < Ni < Fe, was observed of temperature, in which the CO conversion began to increase, while Co achieved a high conversion close to its limit (66.7%) at lower temperature than others, as shown in ). This temperature hierarchy of X and the high conversion of Co seemed to be basically followed by X in X2TiSn, X2TiAl, and X2MnSn, as shown in , although the hierarchy of Fe and Ni in X2TiSn was opposite. This indicates that a main active element is X in X2YZ for CO oxidation.

Table 1. Heusler alloy catalysts used in the subsections 3.1 and 3.2. Top row: X, left column: Y, and cell: Z, in X2YZ.

Figure 4. CO conversion in CO oxidation using [1.2%CO/0.4%O2/He balance] reactant during a continuous heating, for (a) pure X powders, (b) X2TiSn, (c) X2TiAl, and (d) X2MnSn. Colors and symbols in (b–d) correspond to those in (a). Horizontal line at 66.7% is a guide for describing the maximum conversion calculated from the CO:O2 ratio. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

Figure 4. CO conversion in CO oxidation using [1.2%CO/0.4%O2/He balance] reactant during a continuous heating, for (a) pure X powders, (b) X2TiSn, (c) X2TiAl, and (d) X2MnSn. Colors and symbols in (b–d) correspond to those in (a). Horizontal line at 66.7% is a guide for describing the maximum conversion calculated from the CO:O2 ratio. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

shows CO conversions during cycles of heating and cooling between 80 and 600 °C for catalysts with a total surface area of 0.027 m2. Most Heusler alloys exhibit a hysteresis as Fe2TiSn and Ni2TiSn, as shown in . This was due to the irreversible oxidation of catalysts, which was revealed by estimating an excess consumption of O2 from a consumption ratio of CO and O2. In contrast, Co2TiSn showed a small hysteresis, as shown in ). This was due to high oxidation resistance, which was indicated by a small excess O2 consumption and by a tiny XRD peak of oxide even after the reaction under O2-rich conditions. These results indicate that we can control an activity by choosing X and a durability by Y and Z.

Figure 5. CO conversion in CO oxidation using [1.2%CO/0.4%O2/He balance] reactant during heating and cooling cycles, for (a) Fe2TiSn, (b) Ni2TiSn, and (c) Co2TiSn. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

Figure 5. CO conversion in CO oxidation using [1.2%CO/0.4%O2/He balance] reactant during heating and cooling cycles, for (a) Fe2TiSn, (b) Ni2TiSn, and (c) Co2TiSn. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

3.2. Hydrogenation of propyne [Citation11,Citation12]

The alloys listed in were investigated for the hydrogenation of propyne (C3H4) using a reactant of [1%C3H4/55%H2/He balance]. shows conversions of C3H4 for catalysts with a total surface area of 0.027 m2. Most Heusler alloys showed a too small conversion, as well as Co2TiSn. This small conversion was indicated due to residual surface oxides and/or segregated Sn at the surface, because most alloys contain elements having very stable oxides such as Al, Si, and Ti, and because Sn atoms tend to segregate at the surface [Citation11,Citation12]. Co2FeGe, Co2MnGe, and pure Co showed a certain conversion, as shown in , whereas pure Fe, Mn, and Ge revealed almost no activity. This indicates that a main active element was X for C3H4 hydrogenation as well as for CO oxidation. The conversion for Co2FeGe (50%Co) was larger than that for pure Co (100%Co). This indicates that a new electronic structure formed by the alloying of Co, Fe, and Ge was more suited for the C3H4 hydrogenation than pure Co. These results indicate that we can control an activity by choosing not only the main active element, X, but also sub-elements, Y and Z, through the creation of new electronic structures.

Figure 6. C3H4 conversion in C3H4 hydrogenation using [1%C3H4/55%H2/He balance] reactant. The data was obtained during the second cooling cycle of continuous heating and cooling, in which a temporal change in conversion was well settled. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

Figure 6. C3H4 conversion in C3H4 hydrogenation using [1%C3H4/55%H2/He balance] reactant. The data was obtained during the second cooling cycle of continuous heating and cooling, in which a temporal change in conversion was well settled. After Ref [Citation11], reproduced with permission by ACS, https://pubs.acs.org/doi/10.1021/acsomega.6b00299. Further permissions related to the material excerpted should be directed to the ACS.

3.3. Selective hydrogenation of alkynes [Citation5]

3.3.1. High alkene selectivity

The investigation in Section 3.2 indicated that only Co2MnGe and Co2FeGe had an activity for the hydrogenation of hydrocarbons in the alloys listed in . Thus, we further investigated them for selective hydrogenation of alkynes, focusing on selectivity. This reaction is important for removing alkyne impurities from alkene feedstocks because the impurities inhibit the polymerization process of alkenes [Citation16,Citation17]. In industry, Pd-based catalysts have been used for selective hydrogenations of ethyne (C2H2) in ethene (C2H4) and propyne (C3H4) in propene (C3H6) [Citation16Citation18]. We found that non-noble metal catalysts, Co2MnGe and Co2FeGe, had a high alkene selectivity with these reactions.

) present results on the C3H4 conversion and the C3H6 selectivity for C3H4 hydrogenation using a reactant of [0.1%C3H4/40%H2/He balance]. For Co2FeGa0.75Ge0.25 in ), the selectivity was high when the conversion was low, decreasing as the conversion increased, which is typical behavior for ordinary catalysts because of the strong adsorption of alkynes inhibiting re-adsorption of alkenes [Citation19Citation21]. In contrast, Co2MnGe and Co2FeGe showed a high selectivity even when conversion was at 100%, as shown in . show a C3H6 conversion for C3H6 hydrogenation without C3H4 using a reactant of [0.1%C3H6/40%H2/He balance]. Co2FeGa0.75Ge0.25 converted C3H6 to C3H8 in 100% even at room temperature, while Co2MnGe and Co2FeGe showed only a small conversion in the whole temperature. Thus, Co2MnGe and Co2FeGe do not have the ability to hydrogenate alkenes to alkanes, which can be called ‘intrinsic selectivity,’ unlike Pd-based catalysts that need precise controls of reaction conditions to ensure the selectivity [Citation16,Citation17].

Figure 7. (a–c) C3H4 conversion and C3H6 selectivity in C3H4 hydrogenation using [0.1%C3H4/40%H2/He balance] reactant, and (d–f) C3H6 conversion in C3H6 hydrogenation using [0.1%C3H6/40%H2/He balance] reactant, for (a,d) Co2MnGe, (b,e) Co2FeGe, and (c,f) Co2FeGa0.75Ge0.25 catalysts. The figure was remade using the data in Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

Figure 7. (a–c) C3H4 conversion and C3H6 selectivity in C3H4 hydrogenation using [0.1%C3H4/40%H2/He balance] reactant, and (d–f) C3H6 conversion in C3H6 hydrogenation using [0.1%C3H6/40%H2/He balance] reactant, for (a,d) Co2MnGe, (b,e) Co2FeGe, and (c,f) Co2FeGa0.75Ge0.25 catalysts. The figure was remade using the data in Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

Owing to this property, they also showed high selectivity even for the hydrogenation of alkyne impurities in alkene feedstocks using a reactant of [0.1%C3H4 or C2H2/10%C3H6 or C2H4/40%H2/He balance], as shown in ), in contrast to the example for Co2FeGa shown in ). An excessive H2 concentration in the reactants, alkyne:H2 = 1:400, also indicated the great selectivity, because the alkyne:H2 ratio is usually 1:10–20 in other reports, including the reports on binary intermetallic compounds [Citation16,Citation20,Citation22,Citation23].

Figure 8. Alkyne conversions and alkene selectivities in alkyne hydrogenations in the presence of alkenes for (a) Co2MnGe, (b) Co2FeGe, and (c) Co2FeGa. The reactant was [0.1%C3H4 or C2H2/10%C3H4 or C2H4/40%H2/He balance]. After Ref [Citation5], © The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

Figure 8. Alkyne conversions and alkene selectivities in alkyne hydrogenations in the presence of alkenes for (a) Co2MnGe, (b) Co2FeGe, and (c) Co2FeGa. The reactant was [0.1%C3H4 or C2H2/10%C3H4 or C2H4/40%H2/He balance]. After Ref [Citation5], © The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

3.3.2. Control of catalytic properties by elemental substitution

We investigated a change in catalytic properties by elemental substitution (Co2MnxFe1–xGayGe1–y). summarizes the effects due to substitutions of Mn for Fe and Ga for Ge in Co2FeGe, using [0.1%C3H4/40%H2/He balance] in reaction. For Mn substitution, the C3H6 selectivity showed no change, while the C3H4 reaction rate increased with Mn substitution, as shown in ). On the other hand, Ga substitution significantly decreased the selectivity, while increasing the reaction rate. In ), we can see a strong correspondence between the apparent activation energy (Ea) estimated from the rate, and a d-band center (εd) estimated from the density-of-state (DOS) curves calculated for bulk states. This means that the change in electronic structures affected the elementary steps of the reaction. Hard X-ray photoelectron spectroscopy confirmed that actual specimens had valence band structures corresponding to the calculated DOSs, which means that both the sample preparations and the calculations were well conducted. Note that the bulk εd can be used for relative comparison among the alloys, with the same structure consisting of similar elements because the change in DOSs, from bulk to surface, originates from the symmetry breaking at the surface [Citation24]. In addition, Mn substitution for Fe in Co2FeGa and Ga substitution for Ge in Co2MnGe showed similar changes in catalytic properties to those in ), which indicates that the effects of Mn-Fe and Ga-Ge substitutions are independent of each other.

Figure 9. Elemental substitution effects on (a) C3H4 reaction rate at 50 °C and C3H6 selectivity at 200 °C, and (b) experimental Ea and calculated εd in Mn substitution (Co2MnxFe1–xGe) and Ga substitution (Co2FeGayGe1–y) for Co2FeGe. The rate is per surface area and relative to that of Co2FeGe. The figure was remade using the data in Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

Figure 9. Elemental substitution effects on (a) C3H4 reaction rate at 50 °C and C3H6 selectivity at 200 °C, and (b) experimental Ea and calculated εd in Mn substitution (Co2MnxFe1–xGe) and Ga substitution (Co2FeGayGe1–y) for Co2FeGe. The rate is per surface area and relative to that of Co2FeGe. The figure was remade using the data in Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

3.3.3. Possible mechanisms of substitution effects

The Mn substitution certainly changed the electronic structures, increasing εd and causing the monotonic decrease in Ea with the Mn composition, but not causing the monotonic increase in the reaction rate, as shown in . We built a hypothesis that this lack of correspondence between Ea and rate was due to the enthalpy-entropy compensation [Citation25]. For H2 adsorption, the increase in εd strengthens the bond between H and catalyst atoms, which decreases (negatively increases) not only the adsorption enthalpy (ΔHH2,ad), but also the adsorption entropy (ΔSH2,ad) due to the reduction of freedom. This phenomenon is the compensation effect [Citation25]. In the Arrhenius equation of the reaction rate, rAexp(–Ea/RT), the apparent frequency factor (A) and the Ea have the relations: A ∝ exp(ΔS/R) and exp(–Ea/RT) ∝ exp(–ΔH/RT), where ΔS and ΔH are entropy and enthalpy changes for related elementary steps. Thus, if a reaction mechanism is fixed with a change in εd, a reduction in Ea with εd should originate from a reduction in ΔH, which diminishes ΔS, and eventually A. We believe that such compensation was the reason why the rate did not increase monotonically with the reduction in Ea. See [Citation5] for details.

However, the increase in the rate due to Ga substitution seemed reasonable. In addition, the selectivity significantly decreased, unlike in Mn substitution. These results can be explained by assuming that Ga atoms contribute to molecular adsorption, but Ge atoms do not. The high selectivity of Pd-based catalysts is believed to originate from small active sites surrounded by poisonous species, which sterically inhibit the adsorption of larger molecules, alkenes, but allow the adsorption of smaller molecules, alkynes [Citation16,Citation17]. ) shows schematic illustrations of (110) surfaces for Co2FeGe, Co2FeGa0.5Ge0.5, and Co2FeGa. These indicate that the size of adsorption sites is small for Co2FeGe, while it is enlarged by Ga substitution, if Ge atoms have no ability for molecular adsorption but Ga atoms do. Thus, Co2FeGe should have a high alkene selectivity, while Ga substitution should decrease the selectivity. In ), we can also see an increase in the number of sites available for adsorption due to Ga substitution, which indicates an increase in the reaction rate. We believe that this hypothesis was the reason for the Ga substitution effects.

Figure 10. (a) Schematic illustrations of (110) surfaces, and (b) local partial DOS (LPDOS) of Ga 4p and Ge 4p bands at Co2FeGa(110) and Co2FeGe(110) surfaces, respectively. The figure was excerpted from Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

Figure 10. (a) Schematic illustrations of (110) surfaces, and (b) local partial DOS (LPDOS) of Ga 4p and Ge 4p bands at Co2FeGa(110) and Co2FeGe(110) surfaces, respectively. The figure was excerpted from Ref [Citation5], ©The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC, http://creativecommons.org/licenses/by-nc/4.0).

We verified the difference of adsorption ability between Ga and Ge. ) shows local partial 4p DOSs of Ga and Ge at (110) surfaces for Co2FeGa and Co2FeGe, respectively. We can see that the Ga band has higher energy than the Ge band, which indicates that Ga atoms have a higher ability for molecular adsorption than Ge atoms. In addition, theoretical calculations have indicated that typical metals gain the adsorption ability through the hybridization of their p-bands with the d-bands of transition metals in intermetallic compounds [Citation26,Citation27].

4. Other research

There is a catalytic study on Cu2MnAl published in 1935 [Citation13]. We did not know the existence of this paper, written in German, until the recent appearance of a patent on Heusler alloy catalysts [Citation14], which mentioned that 1935 work. Hedvall had intensively investigated the relationship between ferromagnetism and catalysis, and reported in this paper that an increase in CO2 yield over Curie temperature was observed in the catalytic reaction of 2CO →  CO2 + C by Cu2MnAl, as well as in the hydrogenations of CO and C2H4 by Ni [Citation13].

The patent was filed by BASF [Citation14], which claimed the supported fine-particle catalysts of Heusler alloys with X = Mn, Fe, Co, Ni, Cu, and Pd; Y = V, Mn, Cu, Ti, and Fe; Z = Al, Si, Ga, Ge, In, Sn, and Sb. Hereafter, we will only mention the experimental data described in the patent. lists the catalysts, which were prepared by impregnation using fumed silica or gamma alumina as a support, and methanol or water as a solvent, followed by annealing under a gas flow of pure H2, 5%H2/N2, or 10%H2/N2. The XRD indicated that most samples consisted of A2 (bcc) main phase and A1 (face-centered cubic, fcc) secondary phase. The formation of L21 and B2 phases could not be monitored because their fingerprint peaks were hidden by the halo peak of amorphous silica. The XRD patterns for Samples 12 and 13 are likely from the A1 single phase. We cannot evaluate the phases of the catalysts supported on the alumina because the XRD patterns were mostly from the alumina. The distribution of particle size was evaluated for Sample 8. ) shows the size distribution for smaller particles (mainly < 400 nm) observed by high-angle annular dark field scanning transmission electron microscopy. The median particle size was evaluated as 86.6 nm in these particles. ) shows the size distribution for larger particles (mainly > 400 nm) observed by backscattered electron images from scanning electron microscopy, and indicates that the size was distributed around 800 nm for the particles of > 400 nm.

Table 2. List of Heusler catalysts with preparation conditions, including phases indicated by XRD and catalytic properties in Ref [Citation14]. ‘Up’ and ‘Down’ for Knoevenagel condensation means that activity was higher or lower than that for SiO2 support without Heusler catalysts, respectively. Phases in samples 14–17 were not determined (n.d.).

Figure 11. Size distribution for (a) smaller particles (mainly < 400 nm) and (b) larger particles (mainly > 400 nm) in Sample 8 listed in . The figure was excerpted from Ref [Citation14]. with slight rearrangement. See Ref [Citation14]. for details of how to estimate the minimum diameter.

Figure 11. Size distribution for (a) smaller particles (mainly < 400 nm) and (b) larger particles (mainly > 400 nm) in Sample 8 listed in Table 2. The figure was excerpted from Ref [Citation14]. with slight rearrangement. See Ref [Citation14]. for details of how to estimate the minimum diameter.

Samples 1–10 (probably also 11) were tested for the Knoevenagel condensation using benzaldehyde and malononitrile. The fumed silica support showed a certain activity. Samples denoted ‘Up’ in showed a higher activity than the pure fumed silica. Especially, Co2FeAl showed the highest activity, while the inventors considered that it was due to the high activity of Al for this reaction. Samples 12–17 were tested for the selective reduction of NO using a reactant of [0.05%NO/0.05%NH3/5%O2/10%H2O/N2 balance]. Samples 12 and 13 showed a moderate activity for reducing NO and produced a certain amount of N2O. Samples 14–17 showed a high activity with almost no production of N2O.

In addition, a thesis exists on the first-principles calculation for NH3 dissociation on Ni2MnGa and Co2CrGe surfaces [Citation15]. The author, Senanayake, considered that a higher DOS at the Fermi energy would result in a lower activation energy for the dissociation, and explored such materials in Heusler alloys. The Co2CrGe surface had a high DOS, lowering the activation energy to 0.862 eV. The Ni2MnGa surface lowered it to only 1.309 eV, while calculations on 105 kinds of Ni-Mn-Ga compositions revealed that the Ni4Mn11Ga surface had a high DOS, lowering it to 0.575 eV. Although the Ni4Mn11Ga can no longer be called ‘Heusler alloy’ and the preparation is probably impossible, the value was lower than 0.73 eV for pure Fe.

5. Summary and future prospects

5.1. Our studies

Catalytic properties of many Heusler alloys were investigated. The results indicated that X is the main active element, and Y and Z have roles that can change the activity, selectivity, and durability. Especially for the selective hydrogenation of alkynes, non-noble metal catalysts with a great selectivity were discovered, based on the feature that there are so many sets of X, Y, and Z. In addition, the systematic control of catalytic properties by elemental substitution was demonstrated based on the feature that substitution is possible with various elements in a wide composition range. In Co2(Mn,Fe)(Ga,Ge) for the selective hydrogenation, the Mn-Fe and Ga-Ge substitutions brought the ligand and ensemble effects, respectively, and they were independent of each other. Thus, we expect that the catalytic properties can be fine-tuned for specific target reactions. Aside from the practical benefit, this tunability would be useful for investigating general mechanisms of catalysis by intermetallic compounds, because the effects originating from electronic structures and surface elements can be investigated independently and systematically under the same crystal structure.

Since many Heusler alloys contain easily oxidizable elements, as shown in , it is difficult to predict which elemental sets are suited for the reactions involving oxygen. Actually, we have also confirmed the metallurgical and microstructural changes due to oxidation in ongoing studies on the steam reforming of methanol, and the dehydrogenation of 2-propanol. However, some alloys show unique properties depending on the elemental set, like the Co2TiSn for CO oxidation, which showed high oxidation resistance. Therefore, we believe that there is great promise in using Heusler alloys for various reactions.

5.2. Beyond our studies

Materials informatics could be relevant when seeking effective Heusler catalysts [Citation28]. This method usually requires a large number of data sets, especially in the application of catalysts where the properties vary strongly depending on experimental processing conditions [Citation29]. Heusler alloys are a well-defined platform where the components and compositions can be changed without structural change. This would decrease the number of descriptors, reducing data sets required. Recently, predictions using materials informatics from only several tens to hundreds of data sets were achieved even for catalysts [Citation30]. Thus, effective Heusler catalysts could be found using the informatic approach with catalytic experiments under the same reaction conditions.

There have only been a few studies on Heusler catalysts, since their potential was only recently discovered. However, certain researchers have considered the application of catalysts as mentioned in Section 4, and we succeeded in spotlighting the potential of Heusler alloy catalysts. Thus, more researchers will start to study Heusler catalysts. For practical applications, the patent seems to have not succeeded in developing ‘good Heusler catalysts’ because the L21 single phase and a small average particle size (e.g. < 50 nm) with the sharp size distribution were not obtained, and because the properties were not compared to those of practical catalysts.

On the other hand, a certain number of groups have tried to prepare Heusler nanoparticles, however, not for catalytic application. The studies published before 2014 were summarized in the review paper by Felser’s group [Citation31]. L21-Co2FeGa nanoparticles with an average particle size of around 20 nm were successfully synthesized on fumed silica supports by impregnation, although a relatively small amount of extra phases existed in addition to the L21 phase [Citation32Citation35]. The synthesis was also achieved by impregnation into facile carbon nanotubes [Citation36]. The synthesis of L21-Co2FeAl nanoparticles was also reported [Citation37,Citation38]. In addition, Fe2CoGa nanoparticles were synthesized, while the structure was Li2AgSb or Hg2CuTi type, the so-called ‘inverse Heusler’ ()) [Citation39]. To our knowledge, for other L21-Heusler nanoparticles, the size was not small enough, or there was not sufficient information on the size distribution or no proof of the L21 phase, like Co2FeSn [Citation40] and Ni2MnIn [Citation41].

There are still only a few reports of successful, high-quality preparation of Heusler nanoparticles. However, this seems to reflect a scarcity of attempts at synthesis, because the applications have been considered limited. Now, we have a big application for catalysts. Therefore, we believe that many researchers will start developing a variety of Heusler nanoparticles as catalysts.

Acknowledgments

We are grateful to Prof. S. Fujii of Kagoshima University for the electronic structure calculation and to Dr S. Ueda of National Institute for Materials Science for collaborating with us for the electronic structure evaluation by hard X-ray photoelectron spectroscopy, which was performed at BL15XU at SPring-8 under the approval of NIMS Synchrotron X-ray Station (proposal no. 2017B4905).

Disclosure statement

The authors are inventors on a patent application related to this work filed by Tohoku University (JP patent application no. 2017–220616, filed on 16 November 2017).

Additional information

Funding

A part of this work was supported by the Hattori Hokokai Foundation, the Iwatani Naoji Foundation, the Noguchi Institute, and NIMS microstructural characterization platform as a program of ‘Nanotechnology Platform’ (project no. 12024046) of the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan. We also thank the supports from the Program for Creation of Interdisciplinary Research from FRIS, Tohoku University, and from Dynamic Alliance for Open Innovation Bridging Human, Environment and Materials.

References

  • Massalski TB, Okamoto H, Subramanian PR, Kacprzak L, editors. Binary alloy phase diagrams. Vol. 1, 2nd. Materials Park (OH): ASM International; 1990. p. 195–197.
  • Massalski TB, Okamoto H, Subramanian PR, Kacprzak L, editors. Binary alloy phase diagrams. Vol. 2, 2nd. Materials Park (OH): ASM International; 1990. p. 1752–1756.
  • Momma K, Izumi F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J Appl Crystallogr. 2011;44(6):1272–1276.
  • Yin M, Hasier J, Nash P. A review of phase equilibria in Heusler alloy systems containing Fe, Co or Ni. J Mater Sci. 2016;51(1):50–70.
  • Kojima T, Kameoka S, Fujii S, et al. Catalysis-tunable Heusler alloys in selective hydrogenation of alkynes: a new potential for old materials. Sci Adv. 2018;4(10):eaat6063.
  • Graf T, Felser C, Parkin SSP. Heusler compounds: applications in spintronics. In: Yongbing X, Awschalom DD, Nitta J, editors. Handbook of spintronics. Dordrecht: Springer Netherlands; 2016. p. 335–364.
  • Nishino Y, Kato H, Kato M, et al. Effect of off-stoichiometry on the transport properties of the Heusler-type Fe2VAl compound. Phys Rev B. 2001;63(23):233303.
  • Planes A, Mañosa L, Acet M. Magnetocaloric effect and its relation to shape-memory properties in ferromagnetic Heusler alloys. J Phys Condens Matter. 2009;21(23):233201.
  • Heusler F. Magnetic-chemical studies. Verh Dtsch Phys Ges. 1903;5(12): 219.
  • Bradley AJ, Rodgers JW, Bragg WL. The crystal structure of the Heusler alloys. Proceedings of the Royal Society A (Proc. Roy. Soc. A). 1934;144(852):340–359.
  • Kojima T, Kameoka S, Tsai A-P. Heusler alloys: a group of novel catalysts. ACS Omega. 2017;2(1):147–153.
  • Kojima T, Kameoka S, Tsai A-P. Correction to “Heusler alloys: a group of novel catalysts”. ACS Omega. 2018;3(8):9738.
  • Hedvall JA, Ferromagnetische HR. Umwandlung und katalytische Aktivität. IV. Hydrierung von CO und C2H4 über Nickel und CO2-Bildung aus CO über der HEUSLER-Legierung MnAlCu2. Z Physik Chem. 1935;30B(1): 280–288.
  • Mueller U, Sundermann A, Trukhan N, et al., inventor; BASF, SE, assignee. Ternary intermetallic compound catalyst. United States patent US 2018/0243691 A1. 2018 Aug 30.
  • Senanayake NM Exploring Heusler alloys as catalysts for ammonia dissociation [master’s thesis]. Bowling Green (OH): Bowling Green State University; 2016.
  • Borodziński A, Bond GC. Selective hydrogenation of ethyne in ethene‐rich streams on palladium catalysts. Part 1. Effect of changes to the catalyst during reaction. Catal Rev. 2006;48(2):91–144.
  • Borodziński A, Bond GC. Selective hydrogenation of ethyne in ethene‐rich streams on palladium catalysts, part 2: steady‐state kinetics and effects of palladium particle size, carbon monoxide, and promoters. Catal Rev. 2008;50(3):379–469.
  • Samimi F, Modarresi ZK, Dehghani O, et al. Application of response surface methodology for optimization of an industrial methylacetylene and propadiene hydrogenation reactor. J Taiwan Inst Chem Eng. 2015;46:51–64.
  • Yoshida N Catalytic hydrogenation of methylacetylene over group VIII metals: Application of microwave spectroscopy to the analysis of isomeric deuteropropylenes [PhD thesis]. Osaka (Japan): Osaka University; 1971.
  • Studt F, Abild-Pedersen F, Bligaard T, et al. Identification of non-precious metal alloy catalysts for selective hydrogenation of acetylene. Science. 2008;320(5881):1320–1322.
  • Kojima T, Fujieda S, Kato G, et al. Hydrogenation of propyne verifying the harmony in surface and bulk Compositions for Fe-Ni alloy nanoparticles. Mater Trans. 2017;58(5):776–781.
  • Armbrüster M, Kovnir K, Friedrich M, et al. Al13Fe4 as a low-cost alternative for palladium in heterogeneous hydrogenation. Nat Mater. 2012;11:690–693.
  • Liu Y, Liu X, Feng Q, et al. Intermetallic NixMy (M = Ga and Sn) nanocrystals: a non-precious metal catalyst for semi-hydrogenation of alkynes. Adv Mater. 2016;28(23):4747–4754.
  • Tsai AP, Kameoka S, Nozawa K, et al. Intermetallic: a pseudoelement for catalysis. Acc Chem Res. 2017;50(12):2879–2885.
  • Roduner E. Understanding catalysis. Chem Soc Rev. 2014;43(24):8226–8239.
  • Krajčí M, Hafner J. Complex intermetallic compounds as selective hydrogenation catalysts – A case study for the (100) surface of Al13Co4. J Catal. 2011;278(2):200–207.
  • Krajčí M, Hafner J. Intermetallic compounds as selective heterogeneous catalysts: insights from DFT. ChemCatChem. 2016;8(1):34–48.
  • Oliynyk AO, Antono E, Sparks TD, et al. High-throughput machine-learning-driven synthesis of full-Heusler compounds. Chem Mater. 2016;28(20):7324–7331.
  • Takahashi K, Takahashi L, Miyazato I, et al. The rise of catalyst informatics: towards catalyst genomics. Chem Cat Chem. 2019;11(4):1146–1152.
  • Furukawa A, Ikeda T, Okaya T. Materials research method using smart materials informatics. Honda R&D Tech Rev. 2017;29(1):90–97.
  • Wang C, Meyer J, Teichert N, et al. Heusler nanoparticles for spintronics and ferromagnetic shape memory alloys. J Vac Sci Technol B. 2014;32(2):020802.
  • Basit L, Wang C, Jenkins CA, et al. Heusler compounds as ternary intermetallic nanoparticles: Co2FeGa. J Phys D: Appl Phys. 2009;42(8):084018.
  • Wang C, Basit L, Khalavka Y, et al. Probing the size effect of Co2FeGa-SiO2@C nanocomposite particles prepared by a chemical approach. Chem Mater. 2010;22(24):6575–6582.
  • Wang CH, Guo YZ, Casper F, et al. Size correlated long and short range order of ternary Co2FeGa Heusler nanoparticles. Appl Phys Lett. 2010;97(10):103106.
  • Wang C, Casper F, Guo Y, et al. Resolving the phase structure of nonstoichiometric Co2FeGa Heusler nanoparticles. J Appl Phys. 2012;112(12):124314.
  • Gellesch M, Dimitrakopoulou M, Scholz M, et al. Facile nanotube-assisted synthesis of ternary intermetallic nanocrystals of the ferromagnetic Heusler phase Co2FeGa. Cryst Growth Des. 2013;13(7):2707–2710.
  • Pezeshki-Nejad Z, Ramazani A, Alikhanzadeh-Arani S, et al. Influence of the surfactant and annealing rate on the morphology, magnetic and structural characteristics of Co2FeAl nanoparticles. J Magn Magn Mater. 2016;412:243–249.
  • Alikhanzadeh-Arani S, Almasi-Kashi M, Ramazani A, et al. Size effects on the magnetic characteristics of a nanostructured Heusler alloy. J Mater Sci. 2016;51(3):1354–1362.
  • Wang C, Casper F, Gasi T, et al. Structural and magnetic properties of Fe2CoGa Heusler nanoparticles. J Phys D: Appl Phys. 2012;45(29):295001.
  • Li T, Duan J, Yang C, et al. microstructure and magnetic properties of Heusler Co2FeSn nanoparticles. Micro Nano Lett. 2013;8(3):143–146.
  • Aksoy S. Synthesis and characterization of NiMnIn nanoparticles. J Magn Magn Mater. 2015;373:236–239.