Skip to main content
ACS Materials Au logoLink to ACS Materials Au
. 2022 Apr 27;2(4):474–481. doi: 10.1021/acsmaterialsau.2c00016

Trimetallic Pentlandites (Fe,Co,Ni)9S8 for the Electrocatalytical HER in Acidic Media

Mathias Smialkowski , Daniel Siegmund ‡,, Kenta Stier §, Lars Hensgen , Marek P Checinski §,*, Ulf-Peter Apfel †,‡,*
PMCID: PMC9928393  PMID: 36855705

Abstract

graphic file with name mg2c00016_0007.jpg

Recently, pentlandite materials have been shown to exhibit promising properties with respect to the hydrogen evolution reaction (HER). A whole series of trimetallic FeCoNi-pentlandite materials and composites have been synthesized from the elements using high-temperature synthesis and categorized in terms of purity. Furthermore, the electrocatalytic properties regarding the HER were determined and correlated to hydrogen adsorption energies, which were determined by means of density functional theory (DFT) calculations. The relationships between activity and its origin generated in this way help to better understand the pentlandite system and provide meaningful approaches for catalyst synthesis.

Keywords: HER, hydrogen evolution reaction, pentlandite, trimetallic pentlandite, sulfide materials, catalyst poison, DFT

Introduction

Establishing a green hydrogen-based economy is one of the major challenges of the current century to mitigate climate change. Herein, electrolyzers are an elegant way to produce green hydrogen with electrical energy from fluctuating renewables. However, a variety of obstacles must be overcome to master this venture. Among them is the use of efficient electrocatalysts for the hydrogen evolution reaction (HER) under acidic conditions, which should offer a high activity and stability.1 Further, the catalyst should be cheap and environmentally benign.2 Currently, precious platinum is the only industrially used material for this purpose but is of limited availability. Recently, metal-rich transition-metal chalcogenides emerged as promising alternatives for traditional precious-metal-based hydrogen evolution catalysts. In particular, pentlandite materials (Pn) of the general formula M9X8, where M is a transition metal and X is a chalcogenide, have been proven to be potent candidates to replace platinum due to their high HER activity and durability also in the presence of common catalyst poisons such as H2S and CO.3,4 The elemental composition of pentlandites is very flexible and a targeted synthesis of Pn materials with precise stoichiometric control can be achieved by various routes, including hydrothermal, mechanochemical, and solid-state methods.49 Along this line, alteration of catalyst composition allowed for a preliminary optimization of the catalyst HER activity and (Fe,Ni)9S8, (Fe,Co)9S8, (Ni,Co)9S8, and (Fe,Ni)9S8–xSex (x = 1–8) were shown to be suitable lead structures.1013 To find more effective catalysts, we recently synthesized trimetallic pentlandites (Fe,Co,Ni)9S8 with different metal ratios, leading to an application patent for acidic HER.14 While syntheses and applications were subsequently reported in detail, important activity–structure correlations for HER of this class of materials have been discussed only succinctly.15,16 In this study, we probe the accessibility of conceivable trimetallic pentlandite phases and provide a hitherto unparalleled analysis of their catalytic activity in HER in conjunction with density functional theory (DFT) calculations. Thus, we derive conclusive insights into the structure/activity relationship of this catalyst class.

Material Synthesis and Characterization

The trimetallic pentlandites Fe3Co6–xNixS8, (x = 1–5), Fe6–xCo3NixS8, (x = 1–5), Fe6–xCoxNi3S8 (x = 1–5) as well as FeCoNi7S, Fe7CoNiS8 and Fe1.6Co5.6Ni1.8S8, were synthesized following our previously established high-temperature route from the elements.1012,17 The exact experimental conditions and adjustments are provided in the Supporting Information (Table S1). Subsequently, the materials were characterized via differential scanning calorimetry (DSC) and powder X-ray diffraction (PXRD) to confirm the characteristic phase transitions as well as the phase composition. Exemplary DSC and PXRD data of the Co9S8, Fe4.5Ni4.5S8, and Fe3Co3Ni3S8 are shown in Figure 1A,B. In DSC, most compounds show the characteristic pentlandite phase transitions between 600–700 and 800–900 °C, previously described by Sugaki et al., confirming typical Pn-type behavior (Figures S1–S3).9

Figure 1.

Figure 1

Powder diffraction pattern (A) and thermal analysis via DSC (B) of Co9S8, Fe4.5Ni4.5S8, and Fe3Co3Ni3S8. Ternary plot (C) shows Fe–Co–Ni sulfide system including various M9S8-type materials from this and previous studies and gives relation to phase-purity via colorized topology map.10,12 Crystal structure (D) represents the Fe3Co3Ni3S8 pentlandite structure.

However, DSC and PXRD also revealed the monometallic Fe9S8 and Ni9S8 to be mixtures of various iron sulfides or nickel sulfides (Figure S4), respectively, while Co9S8 is the only phase-pure monometallic pentlandite (Figure 1C). This was to be expected, as the pentlandite phase has been shown to be unstable at very high Fe or Ni contents.10,18 In analogy, the bimetallic compounds were synthesized according to previous procedures, by either mechanochemical or high-temperature pathways and exhibit the previously reported PXRD features (Figure S5).4,12 Notably, the trimetallic pentlandite phase is present within very wide limits in the ternary Fe–Co–Ni sulfide system (Figure 1C). These limits are defined by the fact that a high cobalt content leads to rising amounts of impurities (pyrrhotite-like TMS) at a fixed iron content (Fe3Co6–xNixS8, Figure S6A). Furthermore, for compounds with a fixed nickel content (Fe6–xCoxNi3S8, Figure S6C), a rather high content of iron turns out to cause increasing amounts of additional phases, i.e., corresponding nonpentlandite binary metal sulfides, like FeS and NiS. Along this line, compounds with fixed cobalt content, both, with high Fe and high Ni content show equally increasing amounts of extra phases (FexCo3Ni6–xS8, Figure S6B), due to previously explained limitations by Fe or Ni (Figure 1C).10,18 However, these borders are even narrower for trimetallic compounds with three cobalt equivalents, where only Fe/Ni ratios close to 1:1 result in completely phase-pure materials. Despite the observed impurities, the pentlandite phase constitutes the main phase in the majority of trimetallic compounds or can be identified as at least a secondary phase. Furthermore, while cobalt-rich compounds are mainly phase-pure, the iron-rich as well as nickel-rich compounds only contained the pentlandite phase next to several other sulfidic phases (Figures S4–S7) and thus must be considered as composites rather than single-phase materials. This especially becomes important when dealing with their electrochemical behavior, as these can have quite complex and intricate properties.1921

The overall elemental composition of the compounds was additionally investigated via SEM-EDX (Figures S8–S10) and reveals a uniform element distribution within phase-pure pentlandite materials exposing only randomly fractured particles with no defined morphologies with a broad particle size distribution. To further confirm the nature of the compounds, Co9S8, Fe4.5Ni4.5S8, and Fe3Co3Ni3S8 were exemplarily investigated via X-ray photoelectron spectroscopy. The spectra show the expected signals in their typical energy regions concomitant with a surficial oxidation even after Ar+ sputtering (Figures S12–S14).

To gain insight into the material structure, crystals were grown for the trimetallic pentlandite Fe3Co3Ni3S8 via vapor-transport methods.9 The structure (Figure 1D) was determined by single-crystal X-ray diffraction methods (XRD) and is compared with previously obtained structures for the monometallic pentlandite Co9S8 and bimetallic Fe4.5Ni4.5S8.11,2224 The selection allows for complementary insight into the elemental positions and bonding situation within the crystals (Table S2). The data match the previously obtained crystal information from PXRD. The materials form isomorphic structures in the same cubic space group Fmm—as expected for a pentlandite phase. Interestingly, the cell constant significantly shortens with the introduction and increasing amounts of cobalt from Fe4.5Ni4.5S8 over Fe3Co3Ni3S8 to Co9S8. Accordingly, the mean bond lengths change to overall shorter distances, with an exception at the octahedral metal site, where the mean bond length is shortest for Fe3Co3Ni3S8 with 2.376 Å, but longest for Co9S8 with 2.482 Å, while Fe4.5Ni4.5S8 has 2.421 Å (Table S3). This behavior points to a certain preferential occupation of the available atom sites by the different metal, e.g., preferential occupation of Fe at the octahedral metal site.25 Similar observations were made for the tetrahedral coordinated metal in the cluster. Here, the Mt–S bond length slightly decreases from 2.166 Å after the introduction of cobalt to 2.146 Å and slightly increases to 2.149 Å for Co only. As a result, the metal center distances near the position Wyckoff 8C—the previously determined “sulfur depletion site”, which is attributed a key role in the formation of an HER-active Pn-surface—is slightly shortened for the Fe3Co3Ni3S8 (3.505 Å) compared to the Co9S8 (3.510 Å) and Fe4.5Ni4.5S8 (3.537 Å) (Figure S15).3,22 Assuming a similar potential for reductive depletion of sulfur for all pentlandite systems, changes to this position may also result in altered activities for the HER.11,13

Electrocatalytic Hydrogen Evolution

Electrochemical analysis of material pellets was performed in an H-type cell utilizing sulfuric acid as an electrolyte at 25 and 75 °C. In this setup, the cathode and anode were separated by a membrane to avoid cross-contamination of the electrodes by ion-redeposition, which was further excluded by postmortem analysis (Figure S11). The values are listed in Table 1 for selected representatives and in Tables S4–S6 for all compounds made within this study. Tafel analysis yields valuable information about the respective kinetic limitations due to different M–H binding modes and involved catalytic centers (Figures S16 and S17).3,26 A Volmer–Heyrovsky-based mechanism is expected for very high Tafel slopes, like 148 mV dec–1 for Co9S8 and 114 mV dec–1 for Fe3Co3Ni3S8, including the chemisorption of a solvated proton with a subsequent reaction of this species with another solvated proton. In contrast, Fe4.5Ni4.5S8 shows a rather low Tafel slope of 78 mV dec–1, which is assigned to a Volmer–Tafel-type reaction, where the fast chemisorption of protons to the surface is followed by a rate-limiting recombination reaction directly on the catalyst surface. This mechanism also applies for platinum electrode references with a Tafel slope of 46 mV dec–1. Regarding pentlandites, the Tafel slope increases with increasing Co content, while compounds containing Fe or Ni usually exhibit significantly lower values. Thus, a Volmer–Heyrovsky-based mechanism is assumed for Co-rich trimetallic materials (Table S4) and the reaction mechanism shifts to a Volmer–Tafel-type pathway with increasing Fe or Ni content.

Table 1. Key Values η50, as well as Tafel Slopes, Exchange Current Densities log (j0), and CDL Taken from Data at 25 °C (and 75 °C)a.

compound η50 “pre”, mVRHE η50 “post”, mVRHE Tafel “pre”, mV dec–1 Tafel “post”, mV dec–1 CDL “pre”, μF CDL “post”, μF
Co9S8 364 (302) 425 (322) 148 (145) 139 (125) 67.4 (63.9) 48.1 (59.3)
Ni3Co6S8* 405 373 118 88 16.6 12.9
Fe4.5Ni4.5S8 385 379 78 86 10.1 20.1
Fe3Co3Ni3S8 376 (298) 385 (297) 114 (88) 100 (101) 22.1 (42.1) 28.5 (24.2)
Fe3Co2Ni4S8 394 388 102 94 6.9 8.8
Fe3CoNi5S8 383 365 92 85 12.0 14.6
Fe4Co3Ni2S8* 412 388 111 97 47.3 52.3
FeCo5Ni3S8 406 387 113 96 9.3 11.6
Fe2Co4Ni3S8 378 390 108 109 35.8 33.9
Fe1.6Co5.6Ni1.8S8 399 394 108 84 16.3 18.6
a

The values “pre” or “post” 2 h electrolysis at −300 mA cm–2 are listed; “*” marks composite compounds.

To probe the electrochemical HER behavior, we evaluated and compared the overpotentials η50 at −50 mA cm–2 (Figure 2A). The values are within a range of 320–490 mVRHE at 25 °C (Table S5). The values show the interesting progression from high to low iron contents that the HER performance decreases noticeably, with a Ni-only sulfide composite “Ni9S8” being the most active material, while an iron-exclusive sulfide composite “Fe9S8” exhibits the lowest activity. Considering the complex phase composition of the materials, the different reactivities might be attributed to phase-dependent interplays or display decomposition processes. A focused look on the phase-pure compounds, however, reveals that the overpotential decreases when going from high iron or nickel content to a more balanced Fe/Ni distribution (Figure 2B).

Figure 2.

Figure 2

Ternary plots of the Fe–Co–Ni sulfide system. The topology maps give the initial η50 values from LSV (Figure S18) at 25 °C. (A) All materials and composites of the study and (B) only a selection of almost phase-pure compounds.

Furthermore, following the trajectory, higher cobalt contents are preferable to achieve an initially increased HER activity as well. Here, it is noticeable that there is a clear benefit from the incorporation of close to or equal amounts of elements, having Fe3Co3Ni3S8 in the center (376 mVRHE). Likewise, the high Co share in Co9S8 is beneficial as the pure, as-synthesized substance reveals the initially highest activity (364 mVRHE). It is remarkable that upon metal variation, which causes only subtle changes in the materials electronics, the HER activity, as well as the respective reaction mechanism, as shown by Tafel analysis, can be varied with no elaborate surface structuring as needed for other metal chalcogenide catalysts.2729 Thus, to allow for judgment on the variation of active sites due to metal exchange, the electrochemically active surface area (ECSA) was determined (Table S6).3,11,17 In theory, the higher the ECSA—usually represented by the double-layer capacity CDL—the more actives sites are available for the HER. The cobalt pentlandite Co9S8 shows the highest CDL at 25 °C with 67.40 μF, followed by “Ni9S8” (23.54 μF), Fe3Co3Ni3S8 (22.13 μF), Ni3Co6S8 (16.58 μF), Fe4.5Ni4.5S8 (10.11 μF), and Fe3Co6S8 (5.88 μF) with the other compounds showing intermediate values (Figures S20–S25). The activity trend derived from the overpotentials is thus reflected in these measurements. However, a notable exception is the Fe-rich composite “Fe9S8”, which has a very high CDL of 56.24 μF, albeit showing low activity. This again underlines that material alteration and various ad- and desorption processes have a great impact on the activity as well and explain variations observed for other composites measured. An isolated look reveals the phase-pure compounds Co9S8 and Fe3Co3Ni3S8 to follow this trend (Figure 3B) as well.

Figure 3.

Figure 3

(A) Overpotentials before and after 2 h of electrolysis at −300 mA cm–2 and either 25 or 75 °C, respectively, for Co9S8 and Fe3Co3Ni3S8 obtained from LSVs at −50 mA cm–2. (B) Electrochemically active surface area (ECSA) represented by the double-layer capacitance (CDL) obtained from CVs before and after 2 h electrolysis at −300 mA cm–2 at 25 and 75 °C. (C) First derivative of overpotential curves obtained after 2 h electrolysis at −300 mA cm–2.

While the described trend would obviously favor Co9S8 as a better catalyst, for an application, this performance must be kept for an extended period of time with low decay rates at elevated current densities.1 Notably, the trend is well preserved for η10 (Figures S18 and S19) but changes already after a short electrolysis. To fully investigate the catalyst performance, stability measurements are thus key. Therefore, the stability of the materials was subsequently investigated via chronopotentiometry at −300 mA cm–2 for a period of 2 h (Figures S26 and S27) at 25 and 75 °C (Table S5). Different activation and deactivation behaviors can be observed for the materials. Obviously, while being most active in terms of initial LSV measurements, Co9S8 shows a significant deactivation, raising its η50 by over 60 mV (25 °C) and 20 mV (75 °C) as seen in Figure 3A. The evaluation of CDL reveals tolerable changes in the ECSA after long-time experiments (Figure 3B). Moreover, the first derivative of the Co9S8 chronopotentiometry curve perfectly highlights the large changes in the overpotential with time, stressing the low stability of the as-synthesized material during electrolysis (Figure 3C). In contrast, the potential is virtually constant for Fe3Co3Ni3S8 at 25 °C as well as 75 °C, representing a significantly higher degree of stability and rendering it a preferable catalyst material than pure Co9S8. Extended stability measurements of Fe3Co3Ni3S8 for 72 h at −300 mA cm–2 also show good stability of the material (Figure S28). The test implementation of the catalyst in a PEM cell also showed good stability over the experimental period of 10 h (Figure S33).

The materials seemingly undergo a mixed thermal/electrochemical altering process under reductive conditions depending on phase composition and purity. For (Fe,Ni)9S8 type pentlandites, this behavior can partially be attributed to sulfur depletion reactions, a process generally involved for Pn-type materials.22 However, this effect seems less important for Co-rich materials indicated by a largely stable ECSA under reductive conditions as well as SEM-EDX postmortem analysis (Figure S11), thereby suggesting a stronger Co–S interaction in these materials, as shown by the smaller bond length at the depletion sites. It is therefore likely that additional processes, especially in composite materials as well as materials with highly heterogenic Co-containing composition take place, especially in view of the pronounced deactivation behavior of Co9S8 while the ECSA was only moderately affected.

Hydrogen Adsorption Energies

To better assess the described findings on the activity of the various compounds and to understand their origins, the hydrogen adsorption energy ΔEH was calculated for selected compounds. The selection was primarily limited to a few materials that best cover the different ranges of pentlandite-type compositions, namely, the monometallic Co9S8; the bimetallic compounds Fe4.5Ni4.5S8, Ni4.5Co4.5S8, Ni3Co6S8, and Fe3Co6S8; and the trimetallic Fe1.6Co5.6Ni1.8S8, Fe3CoNi5S8, Fe4Co3Ni2S8, and Fe3Co3Ni3S8. Even though some of them were composites, the selection ensured a consistent series in the calculation of ΔEH. The binding energies were calculated via density functional theory (DFT) as a possible descriptor for the catalyst performance during electrochemical hydrogen evolution reaction to facilitate the prediction of novel promising catalyst compositions among pentlandite materials. For this, M9S8-[111]-surfaces with exposed metal sites (thus effectively emulating sulfur depletion sites) have been generated computationally and hydrogen atoms were offered (Figure S29). Examples are shown in Figure 4 for Fe3Co3Ni3S8 and Fe4.5Ni4.5S8. According to Sabatier’s principle, a reactant should neither bind too strong nor too weak to a catalyst surface, to reach optimal performance.30,31 This was the basis for Nørskov et al. to develop a volcano relationship for HER catalysts, using the common logarithm of the exchange current density j0 as a function of the hydrogen adsorption energy ΔEH.32 Log (j0) is considered to give the rate of hydrogen evolution per surface area at a certain potential where the reaction is at equilibrium.

Figure 4.

Figure 4

2 × 2 × 1 Supercell of pentlandite surfaces with adsorbed hydrogen atoms. Fe3Co3Ni3S8 comprises μ2-H, while at the Fe4.5Ni4.5S8 surface, the H is terminally bound. Rhombus marks DFT-calculated unit cell.

They estimated the “top” of the volcano at around −0.24 eV, which has been demonstrated and confirmed by the fact that the potent noble-metal catalysts, e.g., Pd and Pt, were approximately located in that region.

For the herein described catalysts, the adsorption energies are summarized in Table 2. Despite the fact that the calculated surfaces are merely a simplification of the real surfaces (not accounting for different phases, facets, or electrochemical alterations of the surface), initial trends in activity are expected to be derived.33 To account for local differences in metal composition at H-adsorption sites, the possible permutations at each position were included in the calculations. As a first observation, it is important to note that the calculations suggest different initial H-binding modes, caused by compositional alterations, which may even change the overall mechanism of action. In general, terminal μ1-adsorption modes appear less favorable compared to bridging μ2-binding modes leading to overall more positive values of ΔEH, which is expected due to a more stable hydride bonding to two metal sites in μ2–H bonds.

Table 2. H-Adsorption Energies of Different [111]-Pentlandite Surfaces Calculated via DFTa.

surface calc. binding mode ΔEH, eV log (j0), A cm–2 involved metal centers
Ni4.5Co4.5S8* μ1 +0.05 –5.61 Co
Ni3Co6S8* μ1 +0.04 –3.27 Co
Fe3CoNi5S8 μ1 –0.12 –5.45 Fe
Fe4.5Ni4.5S8 μ1 –0.14 –6.09 Fe
Co9S8 μ2 –0.23 –4.15 Co Co
Fe3Co3Ni3S8 μ2 –0.30 –4.67 Fe Co
Fe1.6Co5.6Ni1.8S8 μ2 –0.31 –5.01 Fe Co
Fe4Co3Ni2S8* μ2 –0.68 –5.08 Co Co
Fe3Co6S8* μ2 –0.77 –3.90 Fe Co
a

“*” marks composite compounds.

It is evident from the results that an optimal H-binding to the surface can only be achieved through the interplay of adjacent metal atoms. It is striking that the substitution of Co in Co9S8 for Ni to obtain Ni3Co6S8 leads to an increase in ΔEH while the additional incorporation of Fe (e.g., Fe3Co6S8) further decreases ΔEH. Thus, it appears that either Co-rich or mixed Co–Fe sites are necessary to ensure surface binding in a favorable μ2 mode. However, the presence of Ni likewise seems to modulate the H-adsorption energy even in predominate μ2-binding materials as evidenced, e.g., by Fe3Ni3Co3S8. To shed even more light onto the observations, the metal centers involved in the preferred binding mode need to be examined. It can be observed that μ1–H bonds with Co in principle display higher values for ΔEH compared to hydride bonds to Fe. This is only partly the case for μ2–H bonds as Co9S8 and Fe4Co3Ni2S8 show. Bonds involving homometallic CoCo centers are not necessarily higher in energy than mixed FeCo centers, which highlights the complexity of metal compositions. It is striking that direct bonds to surface nickel are not favored in any way, although it strongly influences the binding properties. The presence of nickel alone ensures that the adsorption energy is raised, e.g., revealed by a comparison of Fe3Co6S8 and Fe3Co3Ni3S8. In numbers, Co9S8 is rather close to the described ideal hydrogen adsorption energy of −0.24 V, with −0.23 eV, while the deviation becomes larger for Fe3Co3Ni3S8 (−0.30 eV)—which is closer to the calculated value of metallic Pt (−0.27 eV)—and Fe4.5Ni4.5S8 (−0.14 eV), while Fe3Co6S8 marks the end with (−0.77 eV). This fits very well with the activity results from the electrochemical measurements.

Cross-Linking Calculations and Experiments

Upon comparison of these theoretical considerations to actual electrochemical measurements of log (j0), it becomes obvious that the activity trend in an isolated view is preserved for at least Co9S8, Fe4.5Ni4.5S8 and Fe3Co3Ni3S8 (Figure 5). The cobalt pentlandite exhibits the highest value for the log exchange current density (log (j0) = −4.15 A cm–2), followed by the trimetallic Fe3Co3Ni3S8 pentlandite (−4.67 A cm–2) and Fe4.5Ni4.5S8 (−6.09 A cm–2), as already assumed from the adsorption energies (Table 2). Setting the calculated ΔEH values for all other compounds in relation to the exchange current density log (j0), possible activity trends may still be derived. However, while the trend is still visible, it is less pronounced in our case. Even though the calculated hydrogen adsorption energy is rather close to the value of platinum metal (−0.27 eV) for Co9S8 and Fe3Co3Ni3S8, the exchange current densities of these catalysts are considerably lower than for PGMs. In contrast, Fe3Co6S8 and Ni3Co6S8 show log (j0) of −3.90 and −3.27 A cm–2, although their hydrogen adsorption energies of −0.77 and 0.04 eV are relatively far from the ideal value. To exclude errors in the experimental methodology, a platinum pellet electrode was measured for comparison. The values of the platinum electrode correspond to those described in the literature.34,35

Figure 5.

Figure 5

Graph showing (A) hydrogen adsorption energy ΔEH, (B) common logarithm of exchange current density log (j0), (C) overpotential η50 at 25 °C (# = value read of at −47 mA cm–2), and (D) Tafel slope of different compounds. The left Y-axis shows compound composition estimation, while the right Y-axis gives the hydrogen coordinating metal centers, both according to DFT. One metal corresponds to a terminally bound hydride, while for two metals, a bridging hydride exists; “*” marks composite materials.

This leads to the conclusion that the correlation of log (j0) and ΔEH is best suited for bare metal electrodes but likely underestimates the complex surface compositions of our sulfidic and pelletized (composite-) materials, with a large number of conceivable active sites. Moreover, the correlation of log (j0) and ΔEH on its own is not sufficient as a descriptor for HER performance but needs additional considerations, like the respective η, as well as exemplarily determination of faradaic resistances to provide a clearer picture of the origin of catalysts activity. For this, the novel compounds were examined in detail for their HER capabilities, as discussed before. Displaying η50 compared to ΔEH (Figure 5) yields a first hint toward a certain correlation between the adsorption of hydrogen on a catalyst surface and its electrochemical HER performance.

The overpotentials tend to decrease for compounds where ΔEH approaches the value −0.24 eV regardless of which metal centers and binding modes are involved, confirming the expected behavior. This trend also shows up during additionally performed galvanostatic electrochemical impedance spectroscopy (gEIS) at −300 mA cm–2 for some selected single-phase compounds (Figures S30–S32). The data show a charge-transfer resistance (RCT) of 4.802 Ω for Co9S8 and 4.867 Ω for Fe3Co3Ni3S8, while the faradaic resistance arising from recombination reactions and electrochemical ad-/desorption (RP) is at 0.749 and 0.625 Ω, respectively (Table S7). For Fe4.5Ni4.5S8, the faradaic resistances RCT and RP are 5.507 and 0.869 Ω in sum higher, explaining the comparably lower performance during the HER. Correlating this again to ΔEH, it becomes evident that just the right binding strength is required for a good performance during the HER. Alongside with the results for the overpotentials, this agrees well with the finding that compounds with a high iron content have a stronger affinity for hydrogen atoms, while nickel has the opposite effect—lowering the reaction rates overall in both cases. The results from the Tafel analysis can also be linked well to the calculations. A Volmer–Heyrovsky-based mechanism was estimated for the examples Co9S8 and Fe3Co3Ni3S8, which favor mainly bridging μ2-hydrides on their active sites during catalysis. Here, seemingly the hydride mobility is a little higher for trimetallic compounds where iron is involved in hydride binding, compared to the pure Co-bound hydride in Co9S8. This could also be reflected in the comparably lower RP value of the trimetallic compound resulting in better mean performance during the HER. In contrast, for Fe4.5Ni4.5S8, a Volmer–Tafel-type reaction path was described, enabled by the terminal binding of hydrides on the Fe sites, which in turn are even more mobile on the catalysts surface, favoring the surface recombination. However, the gEIS data suggest that this process does not result in faster reaction rates, due to the higher faradaic resistances. The deactivation behavior of Co9S8 and also other Co-rich compounds could be attributed to the formation of strongly adsorbed hydrogen species on the surface, probably involving bonding interactions to Co, which hamper the lateral mobility of adsorbed H-species, their recombination on the surface, as well as desorption upon reaction with additional protons, albeit not entirely preventing them. This would be consistent with the calculated results indicating stronger μ2–H bonds on Co-containing surfaces, a less pronounced deactivation at higher temperatures as well as the conservation of ECSA. In total, introducing Co as well as other foreign metals to the Pn-lattice noticeably influences the typical Pn-type HER mechanisms and activity. This behavior can be exploited to tune the electrochemical activity. Furthermore, circumventing S-depletion may provide a useful option in the design of metal chalcogenide catalysts for usage under conditions where a partial release of sulfur has to be avoided, not to mention the preservation of the original catalyst composition.

Conclusions

In summary, we herein demonstrated the preparation of a variety of trimetallic pentlandite compounds (Fe,Co,Ni)9S8 based on iron, cobalt, and nickel. Hydrogen adsorption energies were determined by means of DFT calculations and correlated with various HER performance indicators. Along this line, several activity phenomena could thus be described with good accuracy. It was found that an approximately equal distribution of iron and nickel led to an increase in performance peaking at an equimolar ratio of Fe, Co, and Ni. This observation could be well explained by DFT calculations, as the ideal binding strength of hydrogen to CoCo or FeCo centers focused the adsorption energy in a preferred region of about −0.24 eV as described for the well-known HER catalyst platinum. The mechanism was also decisively altered by the corresponding presence or absence of one of the transition metals. While Co9S8 exhibited a comparatively high Tafel slope (Volmer step with a rate-limiting Heyrovsky step) and Fe4.5Ni4.5S8 exhibited a comparatively low Tafel slope (Volmer step followed by Tafel step), the values of the trimetallic compounds, such as Fe3Co3Ni3S8, were located at a sweet-spot (mainly Volmer–Heyrovsky). From a stability perspective, the phase-pure trimetallic compounds are much more durable than Co9S8, maintaining their activity for a prolonged electrolysis time at −300 mA cm–2. Moreover, the activity of the (Fe,Co,Ni)9S8 did not change over the duration of electrolysis due to the absence of any sulfur depletion mechanism, as is the case with the bimetallic FeNi variants, thus preserving the original composition.

Acknowledgments

The authors gratefully acknowledge funding from Tribotecc GmbH. U.-P.A. is grateful for the financial support by the Deutsche Forschungsgemeinschaft (under Germany ´s Excellence Strategy–EXC-2033–Project Number 390677874) and the Fraunhofer Internal Programs under grant no. Attract 097-602175 as well as CINES. The authors are grateful to Stephan Spöllmann (RUBION) for measuring the XPS data, and Julian Kleinhaus, Tobias Kull, Sebastian Sanden, as well as David Tetzlaff for the helpful discussion and sharing of time and resources. The authors also thank Dr. Andrzej Mikuła (AGH UST Krakow) for the very productive scientific dialog. They especially thank Leonard Messing and Lucas Hoof for their support and provision of the data from the PEM experiments. D.S. gratefully acknowledges funding from the BMBF in the framework NanoMatFutur (“H2Organic”, project number 03XP0421). U.P.A., D.S. and L.H. acknowledge financial support from the BMWK (“CO2-syn”, project number: 03EE5104A).

Glossary

Abbreviations

CP

chrono potentiometry

CV

cyclic voltammetry

DFT

density functional theory

DSC

differential scanning calorimetry

EDX

energy-dispersive X-ray spectrometry

HER

hydrogen evolution reaction

LSV

linear sweep voltammetry

PGM

platinum group metals

PXRD

powder X-ray diffraction

SEM

scanning electron microscopy

XRD

single-crystal X-ray diffraction

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmaterialsau.2c00016.

  • Materials, experimental procedures, and characterization parameters (Section S1); materials synthesis (Section S2); differential scanning calorimetry (Section S3); powder X-ray diffraction (Section S4); scanning electron microscopy/energy-dispersive X-ray spectrometry (Section S5); X-ray photoelectron spectroscopy (Section S6); single-crystal X-ray diffraction (Section S7); Tafel analysis (Section S8); linear sweep voltammetry (Section S9); electrochemical surface area (Section S10); chronopotentiometry curves (Section S11); density functional theory calculations (Section S12); electrochemical impedance spectroscopy (Section S13); and galvanostatic measurement in a PEM setup (Section S14) (PDF)

The authors declare no competing financial interest.

Supplementary Material

mg2c00016_si_001.pdf (3.4MB, pdf)

References

  1. Siegmund D.; Metz S.; Peinecke V.; Warner T. E.; Cremers C.; Grevé A.; Smolinka T.; Segets D.; Apfel U.-P. Crossing the Valley of Death: From Fundamental to Applied Research in Electrolysis. JACS Au 2021, 1, 527–535. 10.1021/jacsau.1c00092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Zhu J.; Hu L.; Zhao P.; Lee L. Y. S.; Wong K.-Y. Recent Advances in Electrocatalytic Hydrogen Evolution Using Nanoparticles. Chem. Rev. 2020, 120, 851–918. 10.1021/acs.chemrev.9b00248. [DOI] [PubMed] [Google Scholar]
  3. Konkena B.; Puring K.; Sinev I.; Piontek S.; Khavryuchenko O.; Duerholt J. P.; Schmid R.; Tueysuez H.; Muhler M.; Schuhmann W.; Apfel U.-P. Pentlandite Rocks as Sustainable and Stable Efficient Electrocatalysts for Hydrogen Generation. Nat. Commun. 2016, 7, 12269 10.1038/ncomms12269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Tetzlaff D.; Pellumbi K.; Baier D. M.; Hoof L.; Shastry Barkur H.; Smialkowski M.; Amin H. M. A.; Grätz S.; Siegmund D.; Borchardt L.; Apfel U.-P. Sustainable and Rapid Preparation of Nanosized Fe/Ni-Pentlandite Particles by Mechanochemistry. Chem. Sci. 2020, 11, 12835–12842. 10.1039/D0SC04525J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Qin W.; Hu B.; Bao D.; Gao P. The Preparation of Co9S8 and CoS2 Nanoparticles by a High Energy Ball-Milling Method and Their Electrochemical Hydrogen Storage Properties. Int. J. Hydrogen Energy 2014, 39, 9300–9306. 10.1016/j.ijhydene.2014.04.061. [DOI] [Google Scholar]
  6. Zhang Y.; Gao Q.; Wang S. Hydrothermal Impregnation Synthesis of Cobalt Pentlandite as Anode Material of H2S SOFC. Ionics 2016, 22, 743–749. 10.1007/s11581-016-1688-z. [DOI] [Google Scholar]
  7. Bezverkhyy I.; Afanasiev P.; Danot M. Preparation of Highly Dispersed Pentlandites (M,M′)9S8 (M, M′ = Fe, Co, Ni) and Their Catalytic Properties in Hydrodesulfurization. J. Phys. Chem. B 2004, 108, 7709–7715. 10.1021/jp049692u. [DOI] [Google Scholar]
  8. Bezverkhyy I.; Danot M.; Afanasiev P. New Low-Temperature Preparations of Some Simple and Mixed Co and Ni Dispersed Sulfides and Their Chemical Behavior in Reducing Atmosphere. Inorg. Chem. 2003, 42, 1764–1768. 10.1021/ic026024j. [DOI] [PubMed] [Google Scholar]
  9. Sugaki A.; Kitakaze A. High Form of Pentlandite and Its Thermal Stability. Am. Mineral. 1998, 83, 133–140. 10.2138/am-1998-1-213. [DOI] [Google Scholar]
  10. Piontek S.; Andronescu C.; Zaichenko A.; Konkena B.; junge Puring K.; Marler B.; Antoni H.; Sinev I.; Muhler M.; Mollenhauer D.; Roldan Cuenya B.; Schuhmann W.; Apfel U.-P. Influence of the Fe:Ni Ratio and Reaction Temperature on the Efficiency of (FexNi1–x)9S8 Electrocatalysts Applied in the Hydrogen Evolution Reaction. ACS Catal. 2018, 8, 987–996. 10.1021/acscatal.7b02617. [DOI] [Google Scholar]
  11. Smialkowski M.; Siegmund D.; Pellumbi K.; Hensgen L.; Antoni H.; Muhler M.; Apfel U.-P. Seleno-Analogues of Pentlandites (Fe4.5Ni4.5S8–YSeY, Y = 1–6): Tuning Bulk Fe/Ni Sulphoselenides for Hydrogen Evolution. Chem. Commun. 2019, 55, 8792–8795. 10.1039/C9CC01842E. [DOI] [PubMed] [Google Scholar]
  12. Smialkowski M.; Tetzlaff D.; Hensgen L.; Siegmund D.; Apfel U.-P. Fe/Co and Ni/Co-Pentlandite Type Electrocatalysts for the Hydrogen Evolution Reaction. Chin. J. Catal. 2021, 42, 1360–1369. 10.1016/S1872-2067(20)63682-8. [DOI] [Google Scholar]
  13. Pellumbi K.; Smialkowski M.; Siegmund D.; Apfel U.-P. Enhancing the CO2 Electroreduction of Fe/Ni-Pentlandite Catalysts by S/Se Exchange. Chem. – Eur. J. 2020, 26, 9938–9944. 10.1002/chem.202001289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Apfel U.-P.; Hensgen L.; Smialkowski M.. Use of Sulfidic Compositions. WIPO Patent WO2020169806A12020.
  15. Mikuła A.; Dąbrowa J.; Kusior A.; Mars K.; Lach R.; Kubowicz M. Search for Mid- and High-Entropy Transition-Metal Chalcogenides – Investigating the Pentlandite Structure. Dalton Trans. 2021, 50, 9560–9573. 10.1039/D1DT00794G. [DOI] [PubMed] [Google Scholar]
  16. Kim Y.; Karuppannan M.; Lee D.; Bae H. E.; Luong Q. T.; Kang S. Y.; Sung Y.-E.; Cho Y.-H.; Kwon O. J. (Fe, Ni, Co)9S8@CS Catalyst Decorated on N-Doped Carbon as an Efficient Electrocatalyst for Oxygen Evolution Reaction. Int. J. Energy Res. 2022, 46, 3145–3154. 10.1002/er.7372. [DOI] [Google Scholar]
  17. junge Puring K.; Piontek S.; Smialkowski M.; Burfeind J.; Kaluza S.; Doetsch C.; Apfel U.-P. Simple Methods for the Preparation of Non-Noble Metal Bulk-Electrodes for Electrocatalytic Applications. J. Vis. Exp. 2017, 124, e56087 10.3791/56087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Knop O.; Ibrahim M. A. Chalkogenides of the transition elements: II. Existence of the π phase in the M9S8 section of the system Fe–Co–Ni–S. Can. J. Chem. 1961, 39, 297–317. 10.1139/v61-036. [DOI] [Google Scholar]
  19. Li H.; Qian X.; Xu C.; Huang S.; Zhu C.; Jiang X.; Shao L.; Hou L. Hierarchical Porous Co9S8/Nitrogen-Doped Carbon@MoS2 Polyhedrons as PH Universal Electrocatalysts for Highly Efficient Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2017, 9, 28394–28405. 10.1021/acsami.7b06384. [DOI] [PubMed] [Google Scholar]
  20. Li Y.; Yin Z.; Cui M.; Liu X.; Xiong J.; Chen S.; Ma T. Interface Engineering of Transitional Metal Sulfide–MoS 2 Heterostructure Composites as Effective Electrocatalysts for Water-Splitting. J. Mater. Chem. A 2021, 9, 2070–2092. 10.1039/D0TA10815D. [DOI] [Google Scholar]
  21. Wu J.; Wang X.; Jiang J.; Lin W.; Zhu S.; Sha J.; Ma L.; Zhao N. In-Situ Synthesis of MoS2/Co9S8 Heterostructure for Efficient HER Electrocatalyst. Mater. Lett. 2021, 292, 129621 10.1016/j.matlet.2021.129621. [DOI] [Google Scholar]
  22. Zegkinoglou I.; Zendegani A.; Sinev I.; Kunze S.; Mistry H.; Jeon H. S.; Zhao J.; Hu M. Y.; Alp E. E.; Piontek S.; Smialkowski M.; Apfel U.-P.; Körmann F.; Neugebauer J.; Hickel T.; Roldan Cuenya B. Operando Phonon Studies of the Protonation Mechanism in Highly Active Hydrogen Evolution Reaction Pentlandite Catalysts. J. Am. Chem. Soc. 2017, 139, 14360–14363. 10.1021/jacs.7b07902. [DOI] [PubMed] [Google Scholar]
  23. Bentley C. L.; Andronescu C.; Smialkowski M.; Kang M.; Tarnev T.; Marler B.; Unwin P. R.; Apfel U.-P.; Schuhmann W. Local Surface Structure and Composition Control the Hydrogen Evolution Reaction on Iron Nickel Sulfides. Angew. Chem., Int. Ed. 2018, 57, 4093–4097. 10.1002/anie.201712679. [DOI] [PubMed] [Google Scholar]
  24. Lindqvist M.; Lindqvist D.; Westgren A. The Crystal Structure of Co9S8 and of Pentlandite (Ni,Fe)9S8. Sven. Kem. Tidskr. 1936, 48, 156–160. [Google Scholar]
  25. Prewitt C. T.; Rajamani V. Thermal Expansion of the Pentlandite Structure. Am. Mineral. 1975, 60, 39–48. [Google Scholar]
  26. Thomas J. G. N. Kinetics of Electrolytic Hydrogen Evolution and the Adsorption of Hydrogen by Metals. Trans. Faraday Soc. 1961, 57, 1603–1611. 10.1039/TF9615701603. [DOI] [Google Scholar]
  27. Voiry D.; Yamaguchi H.; Li J.; Silva R.; Alves D. C. B.; Fujita T.; Chen M.; Asefa T.; Shenoy V. B.; Eda G.; Chhowalla M. Enhanced Catalytic Activity in Strained Chemically Exfoliated WS2 Nanosheets for Hydrogen Evolution. Nat. Mater. 2013, 12, 850–855. 10.1038/nmat3700. [DOI] [PubMed] [Google Scholar]
  28. Eady S. C.; Peczonczyk S. L.; Maldonado S.; Lehnert N. Facile Heterogenization of a Cobalt Catalyst via Graphene Adsorption: Robust and Versatile Dihydrogen Production Systems. Chem. Commun. 2014, 50, 8065–8068. 10.1039/C4CC02920H. [DOI] [PubMed] [Google Scholar]
  29. Li Y.; Wang H.; Xie L.; Liang Y.; Hong G.; Dai H. MoS2 Nanoparticles Grown on Graphene: An Advanced Catalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. 10.1021/ja201269b. [DOI] [PubMed] [Google Scholar]
  30. Sabatier P. Hydrogénations et déshydrogénations par catalyse. Ber. Dtsch. Chem. Ges. 1911, 44, 1984–2001. 10.1002/cber.19110440303. [DOI] [Google Scholar]
  31. Zhu C. R.; Gao D.; Ding J.; Chao D.; Wang J. TMD-Based Highly Efficient Electrocatalysts Developed by Combined Computational and Experimental Approaches. Chem. Soc. Rev. 2018, 47, 4332–4356. 10.1039/C7CS00705A. [DOI] [PubMed] [Google Scholar]
  32. Nørskov J. K.; Bligaard T.; Logadottir A.; Kitchin J. R.; Chen J. G.; Pandelov S.; Stimming U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23. 10.1149/1.1856988. [DOI] [Google Scholar]
  33. Chen M.; Smart T. J.; Wang S.; Kou T.; Lin D.; Ping Y.; Li Y. The Coupling of Experiments with Density Functional Theory in the Studies of the Electrochemical Hydrogen Evolution Reaction. J. Mater. Chem. A 2020, 8, 8783–8812. 10.1039/D0TA02549F. [DOI] [Google Scholar]
  34. Nørskov J. K.; Abild-Pedersen F.; Studt F.; Bligaard T. Density Functional Theory in Surface Chemistry and Catalysis. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 937–943. 10.1073/pnas.1006652108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Trasatti S. Work Function, Electronegativity, and Electrochemical Behaviour of Metals. J. Electroanal. Chem. Interfacial Electrochem. 1972, 39, 163–184. 10.1016/S0022-0728(72)80485-6. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

mg2c00016_si_001.pdf (3.4MB, pdf)

Articles from ACS Materials Au are provided here courtesy of American Chemical Society

RESOURCES