DOI:
10.1039/D5RA03059E
(Paper)
RSC Adv., 2025,
15, 26371-26382
One step laser-induced synthesis of a bimetallic iron–cobalt sulfide for efficient solar light driven, Fenton-like and electrochemical catalysis†
Received
30th April 2025
, Accepted 23rd June 2025
First published on 23rd July 2025
Abstract
Pulsed laser irradiation of an equimolar mixture of FeS2 and CoS2 onto a Ta substrate results in the one-step formation of bimetallic iron cobalt sulfide. The use of complementary analytical techniques, such as scanning electron microscopy, Raman spectroscopy, X-ray photoelectron spectroscopy, X-ray diffraction, high-resolution electron microscopy, and electron diffraction, confirmed the presence of nanocrystalline cobaltpentlandite [FeCo8S8] and maghemite [γ-Fe2O3]. The mechanism by which this occurs involves the reactive interaction of laser-ionized Fe, Co, and S species, which subsequently undergo rapid non-equilibrium cooling and deposition. The higher deposition tendency of CoS2 along with iron ions/atoms leads to the formation of FeCo8S8. This proposed mechanism is supported by density functional theory (DFT), which provides a deeper understanding of the higher thermodynamic stability of Fe in Co1−xFexS2 compared with Co in Fe1−xCoxS2. The FeCo8S8-based deposit exhibited enhanced catalytic efficiency for methylene blue daylight-driven and Fenton-like degradation. In contrast, for solar light-driven degradation of sulfamethoxazole and trimethoprim, the FeCo8S8-based deposit does not show enhanced catalytic activity compared to FeS2 and CoS2. Additionally, electrochemical testing of the oxygen evolution reaction (OER) revealed significantly improved performance for the FeCo8S8-based deposit compared to FeS2 and CoS2 individually.
1. Introduction
Currently, nanocrystalline metal sulfides attract great scientific interest due to their tunable optical, physical and chemical properties. Metal sulfides represent semiconducting compounds where the metal cations can be included in mono-, bi- or multi-form.1 Metal monosulfides have received attention over the past few decades owing to their suitable electronic band gap, band position, exposed active sites, and promising catalytic and photocatalytic activities. Compared to metal oxides, metal sulfides in many cases exhibit shallow valence bands2,3 and a smaller effective mass, which allows a strong quantum size effect.4,5 Metal sulfides can be prepared with various sizes, morphologies, and chemical and phase compositions, which allow excellent tunability of the photon response over a broad range of the solar spectrum. Metal sulfides are also used in rechargeable batteries, specifically as electrocatalytic materials suitable for the oxygen evolution reaction (OER) and oxygen reduction reaction (ORR).6 In this field, extensive efforts have been devoted to replacing noble metals with various non-noble materials such as metals,7 metal oxides,8 metal chalcogenides.9 However, the use of these materials as electrode components is still limited by the weak synergy between intrinsic activity, selectivity, and electrical conductivity.10 Metal oxides exhibit good intrinsic catalytic activity; however, their electrical conductivity is not sufficient.11 Metal sulfides exhibit higher conductivity and similar intrinsic activity compared to metal oxides,12 which is caused by the electronegativity and atomic radius differences between oxygen and sulfur atoms. The properties of monometallic sulfides may be further exceeded by those of bimetallic sulfides because of the synergistic effect between metal sites and also due to the changes in bond distance, bond energies, and bond modes.13,14 Thus, bimetallic sulfides offer enhanced characteristics such as optimal band gap, good electronic conductivity, and intrinsic activity, which are desirable for photo- and electrochemical catalysis. In comparison to the monometal sulfides, bimetal sulfides display rich redox chemistry due to the participation of two metals in addition to higher electronic conductivity than their monometal sulfide and oxide counterparts. Consequently, a significant enhancement of the electrochemical redox properties and synergistic effects from two metal ions are reported for bimetal sulfides applied in lithium ion battery electrodes, sodium ion battery electrodes, metal–air battery electrodes, supercapacitors and water splitting devices.14–16 Despite the exceptional application potential of bimetallic sulfides, the phase and structural design of this compound represents a challenge.17 The formation of bimetallic sulfides is more difficult than that of their monometallic counterparts because the crystallization of bimetallic sulfides is strongly influenced by two metal element properties with distinct kinetics and thermodynamic parameters.18 Specifically, bimetallic iron–cobalt sulfide has been shown to exhibit excellent photocatalytic activity (for the photodegradation of methylene green and rhodamine-B)19 and, electrochemical performance in terms of OER activity,20–24 which are enhanced compared to monometallic sulfides. Hierarchical cobalt–iron sulfide anchored on carbon fibers provides practical support for efficient capture of Hg0 and outperforms traditional iron sulfide and cobalt monosulfide.25
Up to now, bimetallic iron–cobalt sulfides, have been synthesized by several steps containing hydrothermal or solvothermal routes: FeCo8S8 and FeCoS2 nanosheets on reduced graphene oxide has been prepared using cobalt(II) acetylacetonate, iron(III) 2,4-pentanedionate, oleylamine, 1-octadecene, oleic acid, and 1-octanethiol precursors;21 Fe-doped CoS nanoplate was prepared via solvothermal reaction with cobalt chloride hexahydrate, ammonium iron(II) sulfate, thioacetamide, triethylenetetramine;20 CoxFe1−xS2 with different Co/Fe atomic ratios was prepared with cobalt nitrate hexahydrate, iron nitrate nonahydrate hydrothermal process and following chemical vapor deposition (CVD) sulfurization treatment;22 iron cobalt transition-metal sulfide (FeCoS) on Ni foam with nanosheet arrays was fabricated via electrodeposition of cobalt nitrate hexahydrate, iron chloride hexahydrate, thiourea solution;23 hollow FeCo2S4 nanotube multi-tripod arrays supported onto carbon cloth was synthetized by hydrothermal procedure of iron nitrate nonahydrate and cobalt nitride hexahydrate followed by sulfidation process using sodium sulfide nonahydrate;24 cobalt–iron sulfide anchored on carbon fibers was prepared using hydrothermal and calcination route with iron nitrate nonahydrate and cobalt chloride hexahydrate and thioacetamide as a sulfur source;25 cobalt–iron sulfide has been formed also from furyl and ferrocenyl based dithiocarbamate ligands.19 Clearly, the current hydrothermal or solvothermal methods for synthesizing iron–cobalt sulfides typically require the use of challenging, less commonly employed, and potentially toxic chemical compounds, as well as a multi-step process that is both time- and energy-consuming. Furthermore, the reported preparation method often involves drying under specific atmospheric conditions, such as inert gas or vacuum, and may even require additional calcination. Our previous investigations, which were conducted systematically, focused on the pulsed laser deposition of individual monometallic iron sulfide26–28 and cobalt sulfide.29 These studies have shown that when these materials are deposited on an unheated copper substrate, reactive deposition occurs, leading to the formation of bimetallic compounds such as Cu5FeS4 and Co2CuS4. However, when these compounds are deposited onto a tantalum substrate, the deposition of the original FeS2 and CoS2 phases is favored.
In this contribution, we report on the pulsed laser ablation of equimolar mixed FeS2 and CoS2 targets, followed by deposition onto a Ta substrate, resulting in the formation of nanostructured FeCo8S8 and γ-Fe2O3 phases. This represents a facile one-step process for the formation of bimetallic sulfides. The experimental data were supported by density functional theory (DFT) calculation to provide a deeper insight into the structural stability and electronic structure of the iron–cobalt sulfide system. To demonstrate the multifunctional catalytic properties of the FeCo8Fe8 phase, the deposit was tested for daylight-driven and Fenton-like catalytic degradation of methylene blue (MB). Additionally, the ability of the FeCo8S8-based film to decompose sulfamethoxazole and trimethoprim under simulated solar light exposure conditions was investigated. Furthermore, the electrocatalytic activity towards oxygen evolution reaction (OER) were assessed. Comparative evaluations were made of the photo-, Fenton-like, and electrocatalytic characteristics of the FeCo8S8-based coating with those of individual monometallic FeS2 and CoS2.
2. Experimental and computational methods
2.1 Laser-induced synthesis
2.1.1 Preparation of target. The FeS2–CoS2 pellet were prepared at 100 atm on a hydraulic press from equimolar mixture of commercially available iron disulfide (FeS2, 99.98%, Aldrich) and cobalt disulfide (CoS2, 99.98%, Aldrich) powders, which were homogenized by mixing in agate bowl under ambient air before pressing. Individual FeS2 and CoS2 pellets were prepared under the same pressing conditions.
2.1.2 Deposition of metal sulfide films. A 3rd harmonic of pulsed Nd:YAG laser with base wavelength 1064 nm (model Q SMART 850) with wavelength 355 nm and energy 180 ± 5 mJ per pulse, pulse duration: 10 ns, with repetition rate 10 Hz, was focused by lens (f = 15 cm) on the spot area of 0.02 cm2 (laser fluence 2.5 J cm−2) was used for ablation of an equimolar mixture of FeS2–CoS2 target. As a substrate tantalum foil (Ta, 99.98%, Aldrich) was used. The highly focused UV laser (355 nm) irradiation of the FeS2–CoS2 pellet results in fast deposition (the glassy chamber becomes dark immediately after the first pulses) and leads to the formation of a homogeneous dark adhesive Fe–Co–S deposit on the Ta substrate. The ablation was accompanied by a significant blue luminescent zone filling the entire space of the reactor chamber, which suggests high ionization of Fe–Co–S–O species. A vacuum (2 × 10−4 Pa) inside a simple tubular Pyrex reactor (70 mL in volume) was achieved using a turbomolecular pump (HiCube, Pfeiffer). The reactor was equipped with borosilicate glass windows. The duration of irradiation was 6 minutes. The target of the FeS2–CoS2 pellet with a diameter 10 mm and height of 4 mm was positioned vertically in the center of the reactor, and the substrate was positioned perpendicular to the Ta target. For each sample deposition, the vacuum chamber was first opened and the clean substrate was placed inside the Pyrex chamber. The reactor was then closed, and the pressure was lowered. After irradiation, the pressure was increased to atmospheric pressure, the chamber was opened, and the sample with the resulting films was taken for examination. Coats based on individual monometallic FeS2 and CoS2 (used as control samples) were deposited under the same conditions.
2.2 Physico-chemical characterization
X-ray diffraction analysis (XRD) of the input powder, irradiated pressed target, and deposit was performed using a Bruker D8 Discover diffractometer equipped with a silicon-strip linear LynxEye detector and a focusing germanium primary monochromator of Johansson type providing Cu Kα1 radiation (λ = 1.54 056 Å). Data for mineral identification were collected in the 2θ range of 5–70° with a step size of 0.016° and a counting time of 1 s at each step, and detector angular opening of 2.935° phase identification was performed using Diffrac.Eva software v7.3 and the ICDD PDF-4/Axiom 2023–2026 database (Bruker AXS GmbH, Karlsruhe, Germany; Release 2025). Semi-quantitative estimation of mineral composition was performed using the reference intensity ratio method implemented in the Diffrac.Eva software. A Scanning Electron Microscope (SEM; Tescan Indusem) equipped with a Bruker XFlash® 5010 energy-dispersive X-ray spectrometer (EDS; resolution 125 eV) was used to image the surface morphology and measure the elemental composition of layers. Measurement was carried out with accelerating voltage 15 kV. Raman spectra were obtained using a DXR Raman microscope with a diode-pumped solid-state laser emitting at 532 nm using high-resolution gratings working in the range of 50–1800 cm−1 and spectral resolution 2 cm−1 FWHM (Full Width Half Maximum). Raman spectra were measured up to 4000 cm−1. Surface elemental analyses were performed using a Kratos ESCA 3400 X-ray photoelectron spectrometer (XPS; Manchester, UK). C 1s, O 1s, Co 2p, S 2p, Fe 2p, and Ta 4f lines were observed. For measurement, a small piece of tantalum substrate with a characterized layer was fastened to a carbon tape and mounted onto a holder. All the spectra were corrected by shifting the main carbon C 1s peak to 284.8 eV. An ion gun was used for surface etching (Ar+ ions, 10 mA current, and 1 kV accelerating potential). The Shirley background was subtracted and the elemental compositions of the layers were calculated from the corresponding areas. Transmission electron microscopy (TEM) analysis (particle size and phase analysis) was carried out with a Transmission Electron Microscope JEM 2200FS (Schottky) from JEOL operated at 200 kV with a CCD Gatan (Digital Micrograph software), in-column Omega energy filter 366 for EFTEM and EELS analysis, STEM mode with HAADF detector and EDS 80 mm2 SDD (Silicon Drift Detector) X-Max detector from Oxford on scraped samples that were subsequently dispersed in ethanol followed by the application of a drop of diluted suspension on a polymer/carbon coated Cu grid. The diffraction patterns were evaluated using the JCPDS-2 and ProcessDiffraction software package.30
2.3 Photocatalytic and electrocatalytic measurements
The Fenton-like and solar-light driven catalytic activity of the Fe–Co–S deposit compared with individual FeS2 and CoS2 deposits was evaluated in terms of MB degradation in aqueous solution in the presence of H2O2 and in the absence of light (Fenton-like process) and under daylight exposure with an intensity of Φ ∼ 700 lx (solar daylight photocatalysis). The MB solution contained 64.2 mmol L−1 H2O2 (H2O2 was present only in the Fenton-like process) and 0.1428 mmol L−1 MB. The total volume of the solution was 3.5 mL. The Ta substrate with the deposit (covered area of 5 × 5 mm) was placed inside the square quartz on the bottom of the reaction cell (base 1 × 1 cm, height 4.5 cm) filled with MB solution, and the depletion of MB was measured every 20 minutes using a DU 730 Beckman Coulter spectrometer. The catalytic activity of the Fe–Co–S deposit was tested also in terms of solar light driven photocatalytic degradation of two selected antibiotics, trimethoprim (TMP) and sulfamethoxazole (SMX). As a source of sun-light, a visible-light sun simulator (100 W, Oriel LCS 100) with an intensity: 100 klx. UV-Vis (ultraviolet-visible) spectra of TMP and SMX depletion were measured using a Shimadzu UV 1800 spectrophotometer with a resolution of 1 nm. UV-Vis spectra were measured for the first 5 h (each hour) and after each 5 hours during the next 24 h. The electrochemical properties were characterized using a Biologic SP-150 electrochemical workstation at room temperature (25 °C). A three-electrode electrochemical flooded cell setup with a graphite rod as the counter electrode and a Hg/HgO electrode (Gaoss Union, Wuhan, China) was used as the reference electrode. CoS2/FeS2/Fe–Co–S film deposited on tantalum sheet was employed as the working electrode. To evaluate the electrocatalytic water oxidation reaction, the cyclic voltammograms were recorded at 10 mV s−1 in the potential range between 0.98 and 1.8 V versus Reversible Hydrogen Electrode (RHE) in an Ar-saturated 1.0 M KOH solution. The electrode stability test was carried out with Fe–Co–S deposit on Ta by chronoamperometry at 1.7 V vs. RHE in a 1.0 M KOH solution (see Fig. S15†). The EIS measurements were recorded in a frequency range from 100 kHz to 0.1 Hz with an amplitude of 5 mV (peak-to-peak) at open circuit potential.
2.4 Crystal structure and DFT calculations
The crystal structures of CoS2 and FeS2 were obtained from Crystallography Open Database.31 For the desired Fe concentration in CoS2 we constructed the 2 × 2 × 2 and 3 × 3 × 3 supercells and substituted Fe at Co occupation at different concentrations in Co1−xFexS2 and vice versa for Fe1−xFCoxS2. For the calculation we used the projector-augmented wave method (PAW)32 implemented in the Vienna Ab initio simulation package (VASP).33,34 The Perdew, Burke, and Ernzerhof generalized gradient approximation (PBE-GGA) was used for the exchange-correlation functional.35 The calculations are performed in several successive steps. We used the VASP code for geometry optimization, electronic structure, and magnetic calculations. All convergence parameters in the code were carefully checked. In the calculation, the 3d74s2 electrons of Co, 3d64s2 electrons of Fe, and 3s23p4 electrons of S were treated as valence electrons. In VASP code the geometries have been relaxed using the conjugate gradient method with forces estimated using the Hellman–Feynman theorem. The energy cut-off was set to 350 eV. A Γ-centered k-point mesh of 4 × 4 × 4 was used for the Brillouin zone sampling. The energy and force convergence criteria were set as 10−6 eV and 10−3 eV Å−1, respectively.
3. Results and discussion
3.1 Spectroscopic characterization of input materials
The XRD pattern of the original CoS2 powder (Fig. S1a†) confirms the expected majority of the cubic cattierite [CoS2] (PDF 04-003-1962) and trace amount Co–Fe sulfates, e.g., cobaltkieserite [CoSO4·H2O] (PDF 01-070-2104; 3.39 Å and 2.50 Å) and mikasaite [Fe2(SO4)3] (PDF 00-033-0679; 6.01 Å and 3.00 Å). FeS2 original powder in agreement with XRD (Fig. S1b†) consists of a major ratio of pyrite [FeS2] (PDF 04-003-1989), and a trace amount of Fe sulfates, e.g., szomolnokite [FeSO4·H2O] (PDF 04-014-9807; 3.44 Å and 3.10 Å) and ferricopiapite [Fe4.67(SO4)6(OH)2·20H2O] (PDF 00-029-0714; 9.06–3.58–3.32 Å). Trace phases were determined based on the d-spacings, which are given in parentheses. Their accurate determination is prevented by the low concentrations in the samples. XRD pattern of pressed mixed target (Fig. S1c†) exhibits phase composition consisting of pyrite (PDF 01-071-5208), rhomboclase [FeH(SO4)2·4H2O] (PDF 00-027-0245) and cobaltkieserite (PDF 00-015-0701). Smaller diffraction intensity of cattierite (PDF 00-041-1471; 3.19–2.76–2.47–2.26–1.95 Å) was also detected. Qualitatively, no phase transformations occurred during the preparation of the target; however, indicated by the line intensities of the individual phases, the ratio of both iron and cobalt sulfates increased compared to the original majority of the FeS2 and CoS2 phases in the individual powders. This observation is consistent with the different form of the measured samples. While powder measurement allows irradiation of entire particles volume and detection of subsurface phases, the hydraulically pressed target exhibits a compact surface with a possible higher ordered orientation, providing stronger diffractions of surface sulfates. In addition, hydraulic pressing causes strong friction at the contact surfaces, which could promote the reaction of surface moisture and sulfides and the formation of a thicker surface layer of sulfates. However, subsequent results show that this surface layer of sulfates is probably removed/decomposed during the first pulses of laser irradiation and does not significantly affect the laser-induced reaction between CoS2 and FeS2.
The X-ray diffraction (XRD) pattern of the target, analyzed before and after laser ablation, exhibited no discernible alterations in phase composition, indicating that no bulk phase transformations occurred in the remaining target material during the processes of pressing and pulsed laser irradiation.
The Raman spectra of the original FeS2 powder (Fig. 1a) exhibits intense sharp bands at 376 and 338 cm−1 and less intense peaks at 424 and 480 cm−1, which corresponds to the pyrite.37 The less intense peak centered around 1085 cm−1 correspond to the szomolnokite,37 which is in agreement with the XRD measurements and it reflects hygroscopic behaviour of the FeS2 powder. The CoS2 powder shows a typical Raman spectra pattern (Fig. 1b) with a dominant band at 665 cm−1 and less intense peaks at 290, 390, 480, 850 and 1050 cm−1.36 Typically, the peaks at 475, 517, and 676 cm−1 are attributed to the CoS phase, while the peaks at 290 and 393 cm−1 belongs to CoS2,36 suggest that the original powder consists of a mixture of both CoS2 and CoS phases. The less intense band centered around 1053 cm−1 could reflect the minor contribution of the hydrate of CoSO4,37 which is in line with XRD data and the hygroscopic nature of CoS2 powder.
 |
| Fig. 1 Raman spectroscopy of original FeS2 powder (a); original CoS2 powder (b); the Fe–Co–S deposit obtained by pulsed laser deposition of mixed FeS2–CoS2 target (c). | |
The Raman spectra of the CoS2–FeS2 target (Fig. S2†) revealed peaks corresponding to the phase compositions of the original FeS2 and CoS2 powders at 1050, 480, 390, 376, 338, and 290 cm−1. Peaks centered at 376, 338 and 424, 480 cm−1 correspond to the pyrite, whereas peaks at 290, 390 and 665 cm−1 are assignable to CoS2.
The appearance of a more intense peak centred at approximately 1050 cm−1 in the region of Co/Fe sulfates and intensity decreasing of 665 cm−1 peak belonging to CoS are in line with XRD data and are commented above.
3.2 Characterization of Fe–Co–S deposit prepared by pulsed laser ablation
3.2.1 Spectroscopic characterization of Fe–Co–S deposit. Typical Raman spectra of the deposit prepared by laser ablation of the mixed CoS2–FeS2 target (Fig. 1c) shows broad peaks centered at 300, 480, and 663 cm−1. This pattern suggests a partially amorphous phase, which could be assigned to both CoS2 and FeS2 (ref. 36 and 37) species and/or their mixed system, considering that the broad peaks overlap peaks assignable to both individual sulfides. The absence of a peak situated around 1050 cm−1 suggests that no significant amount of cobalt or iron sulfates were formed during pulsed laser deposition of the Fe–Co–S coat.X-ray photoelectron spectroscopy measurements were used to determine the surface composition of the Fe–Co–S coat (XPS of the original CoS2 powder is provided in Fig. S3 and S4† for comparison with XPS of Fe–Co–S deposit). The XPS quantification results corresponding to the binding energies (BE) of several core levels (C 1s, O 1s, Co 2p, S 2p and Fe 2p) are summarized in Table 1. No Ta signal was detected in any of the Ta 4f spectra of the samples examined. The most significant core levels spectra of the samples after etching are presented in Fig. 2.
Table 1 Quantification summary of the XPS data in dependence on etch time
Etch time |
Co 2p total [at%] |
S 2p sulfate [at%] |
S 2p sulfide [at%] |
Fe 2p total [at%] |
O 1s C O [at%] |
O 1s metal [at%] |
0 |
3.32 |
3.60 |
16.61 |
2.66 |
23.32 |
2.11 |
60 |
7.23 |
2.15 |
26.83 |
5.14 |
10.94 |
4.73 |
240 |
12.83 |
1.04 |
31.29 |
8.78 |
5.11 |
7.14 |
 |
| Fig. 2 S 2p spectrum (a), Co 2p3/2 spectrum (b), Fe 2p spectrum (c) of Fe–Co–S deposit after etching. | |
The main signal from the C 1s region was ascribed to adventitious carbon, that is, carbon-containing impurities from air. The spectra consisted predominantly of signals ascribed to the C–C bond used for calibration set to 284.8 eV with minor contributions at approximately 287 and 288 eV, which are usually assigned to C–O and C
O, respectively. Spectrum O 1s consisted of broad signal at around 531.7 eV, which can be ascribed to C
O from adventitious carbon and/or surface-bound water or hydroxides (the percentage representation is in the column “O 1s_C
O [at%]” in Table 1). This broad peak might also contain a minor contribution from C–O at approximately 533 eV and 532 eV (only in the surface layer), which is assignable to sulfates. After etching, a peak at approximately 530 eV corresponding to the metal oxide emerged (the percentage representation is in the column “O 1s_metal [at%]” in the Table 1).
The spectra from the S 2p region (Fig. 2a and Table 1 column “S 2p_sulfide [at%]”) showed a broad peak with a tail towards higher binding energies. The S 2p signal is usually fitted by two S 2p3/2 and S 2p1/2 peaks with a 1.16 eV shift between the peaks. The measured spectra contained several contributions. The main peak centered at 161.8 eV was ascribed to sulfide. It is not possible to distinguish between the different sulfides, i.e. between CoS2 and FeS2. Other two contributions centered at 162.9 and 164.5 eV were ascribed to terminal and central sulfur atoms in polysulfides.39 There was also a minor contribution from sulfite at 166.7 eV. Compared to the surface spectra, the spectrum after etching did not contain any sulfate contribution.
The XPS region of Co 2p3/2 (Fig. 2b and Table 1 column “Co 2p [at%]”) showed a sharp maximum at 778.6 eV with a tail towards the higher binding energies and a minor signal corresponding to the satellite structure at approximately 785 eV. Deconvolution of the Co 2p signal is rather complicated. The original CoS2 powder (ESI S1 and S2†) was measured to obtain the best fitting parameters for sulfide. The majority of cobalt sulfide was confirmed by the peak at 778.6 eV. The minor contribution of cobalt oxide suggested a peak at 779.9 eV. The oxide fitting parameters were derived based on detailed literature. During the fitting process, the Co3O4 signal shape was determined as the most likely oxide compound present in the sample, with a peak at 779.9 eV.40 The XPS signal of the Fe–Co–S deposit after etching was fitted by sulfide and oxide fitting patterns, showing a clear majority of sulfidic cobalt and less than 10 rel.% of cobalt oxide. The Fe 2p spectrum (Fig. 2c and Table 1 column “Fe 2p [at%]”) did not show any sign of a sulfide peak pattern. Only the Fe2O3 pattern was observed on the surface. After the etching, Fe2O3 pattern with two overlapping Fe 2p3/2 peaks at 708.6 and 711.0 eV together with satellite contribution at 730.5 pointed out Fe2+ enrichment which is, however, normal for Fe2O3 samples undergoing etching process.40 In summary, XPS analysis provides compelling information that the Fe–Co–S deposit exhibits sulphates contributions only in the superficial layer while after etching, only sulfidic sulfur has been detected. Interestingly, cobalt (or bimetallic Fe–Co) sulfide is a dominant sulfide compound, in contrast to iron, which appears to be completely oxidized.
X-ray diffraction of the Fe–Co–S deposit (Fig. 3) predominantly shows diffraction lines of cubic tantalum originating from the substrate. However, upon closer inspection of the XRD spectra, cobaltpentlandite [FeCo8S8] (PDF 04-005-0440; 2.99–2.86–2.28–1.91–1.75 Å)63,64 was also detected (together with minor signals attributed to monoclinic hydrated Co sulfates, marked as m-cobalt sulfates in Fig. 3).
 |
| Fig. 3 XRD patterns of Fe–Co–S deposit on Ta. | |
3.2.2 Microscopic characterization of Fe–Co–S deposit. SEM images (Fig. 4a and b) show predominantly round particles with sizes of approximately 1 μm and sub-μm. Some of the particles exhibited interesting perforated and/or hollow morphologies. These features are in accordance with vaporized and plasma-produced clusters ejected from the target surface and quenched upon deposition as already observed and investigated in previous related studies.26,27,29 Moreover, identical hollow objects have been observed for pulsed laser deposition of hygroscopic compounds,38 when laser-induced evaporation and expansion of surface-absorbed moisture of ejected species could result in subsequent solidification of deposited material in the form of hollow morphologies. This assumption is in accordance with here observed hygroscopic behavior of the deposited material (see Section 3.1). The round-shaped objects indicate the rapid cooling of the solidifying gas phase/liquid droplets with unheated Ta surface, which may occur in a metastable state.26,27,29 EDS analyses indicated the presence of Fe, Co, S, Ta, and O. The average atomic ratio of the film on Ta was Fe0.17Co0.18S 0.42O0.23. These values are in line with the partial oxidation of FeS2 and/or CoS2 and with approximately 1
:
1 ratio of Fe/Co and 1
:
2 of Fe/S and Co/S ratios, which corresponds to the chemical composition of the original irradiated target. Elemental mapping (Fig. 4c–f) depicts the homogenous distribution of tracked elements with sub-micro regions of higher concentration, which indicates the distribution of the particles.
 |
| Fig. 4 SEM images of the Fe–Co–S deposits on Ta (a) and (b) and its elemental mapping (c)–(f). | |
TEM images of the Fe–Co–S coat provide more detailed insight into the morphology of the deposit obtained by laser ablation of the FeS2–CoS2 target. Fig. 5a and b shows irregular agglomerates in the size range of hundreds of micrometers, which consist of smaller spherical nanoparticles sized around tens and units of nanometers, which are embedded in an amorphous matrix. High-resolution electron microscopy image analyses revealed nanocrystalline regions whose interlayer spacing d = 0.228 (331) fits with the cubic cobaltpentlandite (PDF 04-005-0440)63,64 (Fig. 5c), and the nanoobjects with crystalline spacing d = 0.321 nm (205) corresponding to the tetragonal maghemite [γ-Fe2O3] (PDF 015-0615)65 (Fig. 5d). These TEM and HRTEM data were in agreement with the Raman and X-ray spectroscopies.
 |
| Fig. 5 TEM images of Fe–Co–S particles (a) and (b) and HRTEM images depicting interlayer spacing of FeCo8S8 (c) and γ-Fe2O3 (d). | |
3.3 Plausible mechanism of FeCo8S8 phase formation induced by pulsed laser ablation
To propose the mechanism of the bimetallic sulfide FeCo8S8 phase we should consider (i) the original structure of the CoS2–FeS2 target, (ii) the processes occurring during pulsed laser deposition in a gaseous plume, and (iii) the cooling collision of excited ejected species with an unheated substrate. Raman spectroscopy and XRD revealed that the pristine and irradiated targets consisted of pyrite, cattierite, rhomboclase and cobaltkieserite. During pulsed laser ablation the irradiated phases undergoes chemical, phase and structural transformations, considering that the final Fe–Co–S deposit composed of partially amorphous phase and partially crystalline nano-entities where complementary analyses revealed the presence of cobaltpentlandite, maghemite, minor amount of cobalt oxide and only very superficial contribution of cobalt sulfates. Thermodynamic non-equilibrium high-temperature reactions that could occur during pulsed laser ablation on the irradiated target, which occur via interaction of ejected highly energized ionized species in the gas phase are favorable for the deposition of materials with a morphology and phase composition that is different from their bulk progenitors. Thus, the formation of cobaltpentlandite along maghemite, a minor amount of cobalt oxide, and an amorphous phase could be attributed to the gaseous interaction of the ejected excited Fe, Co, and S species, which subsequently collided with the unheated Ta substrate. The formation and interaction of inorganic cluster ions in the plume generated by laser ablation have been thoroughly studied, for example, for metal oxides, phosphides, chalcogenides, and metal carbonyls.41–45 Our recent studies described pulsed laser ablation of individual FeS2 (ref. 26 and 27) and CoS2 (ref. 29) on a Ta substrate (considered as an inert substrate), where the deposition of the original pyrite and cattierite (cubic CoS2 phase) took place. Interestingly, in the case of pulsed laser irradiation of the FeS2–CoS2 mixture, complementary analysis confirmed that the resulting deposit did not contain the FeS2 phase but was composed predominantly of cobaltpentlandite along the maghemite. It is questionable whether iron was detected only in the form of its oxide and incorporated into the cobaltpentlandite. Owing to the lack of theoretical models and calculations for such highly thermodynamically non-equilibrium states, we can only state the final products of the reaction and discuss the possible kinetics of a given event. The main factors that could represent the clue for the explanation of given mechanism are proposed to be higher decomposition temperature of CoS2 at 650 °C compared to decomposition temperature at 550 °C of FeS2,46,47 whereas the chemical affinity of Co to S seems to be higher than between Fe and S.47 Considering ∼100 °C higher thermal stability of CoS2 compared to FeS2 we can propose that during collision of excited gaseous Fe, Co, S species with unheated Ta substrate, Co and S in the given ratio could exhibit tendency for deposition of CoS2 at higher temperature compared to Fe and S to create FeS2 phase. The former desublimation of CoS2 in the presence of iron species could result in the incorporation of iron atoms into CoS2 and the formation of bimetallic cobaltpentlandite, while residual iron from still decomposed FeS2 (due to disruption of the original stoichiometry) did not re-bond to sulfur, but was deposited in an atomic form, which was subsequently oxidized.
Simplistic scheme of proposed mechanism is given in Fig. 6.
 |
| Fig. 6 Simplified scheme of the proposed kinetics of laser induced FeCo8S8 and Fe2O3 reactive deposition: pulsed laser irradiation of FeS2 and CoS2 mixture target (1), laser-induced formation of ionized excited Fe, Co, S species interacting in the gas phase (2), colliding of excited Fe, Co, S species with unheated Ta substrate (3), former tendency of CoS2 for the deposition in the presence of Fe excited species which results in inclusion of Fe atoms into depositing CoS2 (4), formation of FeCo8S8 and deposition of atomic iron, which oxidized into Fe2O3 (5). | |
The final step (4 to 5 in Fig. 6) of above proposed reactive process is very rapid. Optical emission spectroscopy (OES) of plasma induced by ablation of pure iron sulfide performed under related conditions revealed the plasma emission with a lifetime of around 500 ns.46 The collision of excited ionic species with unheated Ta surface inevitably results in acceleration of plasma cooling which leads us to assumption that the reaction which took place is concerted. The speed of the process acts in a suppressive manner towards the formation of eventual intermediates, however in the previous step (step 3: intermixing of highly excited species in the plasma state) a transient may be formed. This assumption is supported by previous studies on pulsed laser-induced reaction of gaseous reactants (1,3-disilacyclobutane and dimethyl selenide) where LIF (Laser Induced Fluorescence) experiment allows detection of transient gaseous SiGe, whose radiative lifetimes was assigned 17–20 ns.66 Deeper insight into the direct connection between transient phenomena in the plasma state and the final structure formed by the de-sublimation of plasma species onto the substrate will require detailed OES and LIF experiments.
3.4 Computational analysis
In order to achieve insight into the thermodynamic stability and an electronic structure of examined phases and thus clarified more deeply mechanism of FeCo8S8, DFT computational analysis have been performed and described below.
3.4.1 Formation energy. In DFT, formation energy is an important parameter that plays a key role in understanding the relative stability of different atomic substitutions in crystal structures and chemical reactions. The formation energies of Co0.99Fe0.01S2 and Fe0.99Co0.01S2 were calculated according to eqn (1) and (2): |
 | (1) |
|
 | (2) |
where E(Co1−xFexS2) is the total ground state energy of Co1−xFexS2 and E(Co), E(Fe), and E(S) are the total ground state energies of individual Co, Fe, and S atoms, respectively, in their standard configuration. The calculated formation energy of Co1−xFexS2 per atom was −0.10138 eV where in Fe0.99Co0.01S2 the calculated formation energy was −0.01364 eV.
3.4.2 Magnetic properties and electronic structure. To explore the magnetic nature of CoS2, Co1−xFexS2 and FeS2, in DFT, the total energies from the ground state calculations were compared for different magnetic states, as shown in Table 2. The magnetic state can be represented according to the orientation of the spin configurations in the ferromagnetic, antiferromagnetic or paramagnetic states. The configuration with the lowest total energy corresponds to the ground state of the FeS2‚ CoS2, Fe1−xCoxS2 and Co1−xFexS2 that provide insights into its magnetic properties. The energy difference between FeS2 and Fe1−xCoxS2 is approximately 0.2 meV likely to exhibit a paramagnetic state, while CoS2 and Co1−xFexS2 are ferromagnetic at 0 K with a half-metallic nature. It is characteristic of Co1−xFexS2 that density of states (DOS) has a dominant contribution at the Fermi energy constructed only from the spin-up band, while the spin-down state shows a gap revealing a half-metallic spin state (see Fig. 7a). The localized kinks in the gap of the spin-up DOS formed by the Fe 3d state are slightly shifted to higher energies from Co0.99Fe0.01S2 to Co0.969Fe0.031S2 but the effect on the band gap in spin-down channel is negligible. The magnetic moment is produced by the Co-3d and Fe-3d states as Co = 0.889 μB, Fe = 0.266 μB and S = 0.031 μB. The magnetic moment at the S site was mainly induced by the Co-3d electron. According to the literature the Curie temperature of CoS2 is 122 K.48
Table 2 Ground state energy of paramagnetic state minus ground state energies of different magnetic configuration
Spin configuration |
Ground states energies in (eV) |
FeS2 |
CoS2 |
Fe0.99Co0.01S2 |
Co0.99Fe0.01S2 |
Paramagnetic |
0.0000 |
0.0000 |
0.00000 |
0.0000 |
Ferromagnetic |
0.0002 |
0.8340 |
−0.00001 |
−4.0904 |
Antiferromagnetic |
0.0005 |
0.2683 |
−0.00013 |
−0.9861 |
 |
| Fig. 7 Effect of Fe doping on DOS of Co1−xFexS2 with different concentration (a), calculated total density of states of Co1−xFexS2 (x = 0.009, 0.031) in the paramagnetic state (b), angular-momentum-projected density of states of Co1−xFexS2 (x = 0.031) in the paramagnetic state (c). | |
Based on Curie temperature of CoS2 the electronic structure calculation of Co1−xFexS2 were performed in paramagnetic phase. Fig. 7b shows the paramagnetic total DOS of Co1−xFexS2 (x = 0.009 and 0.031). Increasing the Fe concentration in CoS2 slightly shifts the bandwidth (−2.2 to −1.0 eV), whereas the shift below the Fermi energy (EF) is hardly visible. This is because the contribution of the Fe 3d state is very small around the Fermi energy, as shown by the angular-momentum-projected density of states (PDOS) of Co1−xFexS2 (x = 0.031). The DOS around EF is mainly constructed from Co-3d states and S-3p states and Fe-3d states do not influence the Co-3d and S-3p state (see Fig. 7c). The total DOS Fe0.99Co0.01S2 show semiconducting nature having the band gap of 1.3 eV (see Fig. S5 in ESI†).
3.4.3 Discussion of computational analysis. The aim of DFT part was to obtain reliable quantitative information regarding the structural stability and electronic structure. Based on the calculated formation energies a deeper understanding of the thermodynamic stability of Fe in Co1−xFexS2 is possible, while Co in Fe1−xCoxS2 is less stable. Therefore, our theoretical prediction provides evidence for why the pulsed laser irradiation of the FeS2–CoS2 mixture does not contain the FeS2 phase but is mainly composed of cobaltpentlandite along with the maghemite (see Section 3.2). The Fe bonded with O is thermodynamically more stable than that bonded with S. Additionally, the electronegativity difference of Fe–O is greater than that of Fe–S, which also favors Fe–O ionic bonding in FeS2–CoS2 mixture. The increasing Fe concentration in Co1−xFexS2 was carefully checked (see Fig. S6 in ESI†) which intensifies the Fe-3d from localized to delocalized and merged with a bandwidth from −1.2 to −2.2 eV but the DOS around EF is unchanged which provide a clue that at equilibrium Fe concentration (up to 10%) does not change the electronic structure significantly and Co1−xS2 stay half-metallic. The photocatalytic measurement of Co1−xFexS2 is close to that of CoS2 (half-metallic) compared to FeS2 (semiconductor), which shows that the present DFT calculations are in agreement with the experimental results. To demonstrate the application potential of the achieved Fe–Co–S deposit, photo-, Fenton-like and electrochemical catalytic tests were performed and are described in the following sections.
3.5 Catalytic and electrochemical performance of Fe–Co–S deposit
The catalytic activity of the Fe–Co–S deposit on the Ta substrate was examined by MB degradation in the terms of Fenton-like process and the photocatalytic process (under daylight exposure) in comparison with the single FeS2 and CoS2 deposits. MB dye was selected for primary catalytic tests as the model standard dye in aqueous solution to demonstrate the primary degradation ability of the organic compounds.
3.5.1 Fenton-like catalytic activity for MB degradation. After immersion of the examined deposits into MB solutions, small bubbles were generated on the surfaces, suggesting an ongoing Fenton-like degradation process. The temporal evolution of the MB degradation measured by UV-Vis spectrometer (see the conditions in experimental) for Fe–Co–S compared to individual FeS2 (ref. 26 and 27) and CoS2 (ref. 29) deposits for the Fenton-like process and solar light driven decomposition is depicted in Fig. 8a and b. The MB decomposition in terms of the Fenton-like process was monitored for 3 hours (180 minutes), whereas the spectra starts to be recorded after 20 minutes in order to filter out the effect of MB adsorption onto the Fe–Co–S coat (Fig. 8a). The degradation of MB after 180 minutes corresponds to residual relative concentration of 82% for FeS2 deposit, 40% for CoS2 deposit and 30% for Fe–Co–S deposit. The depletion of MB reflects MB degradation owing to the absence of an adsorption peak at 256 nm, assignable to the leuco-MB form. UV-Vis spectra of MB depletion are available in the supplemental data (Fig. S7† – for FeS2; Fig. S8† – for Fe–Co–S; in supplemental data in ref. 29 for CoS2). The activities of the FeS2 and CoS2 phases for Fenton-like degradation processes have been described in previous studies.29,49 To the best of our knowledge, the Fenton-like activity of bimetallic FeCo8S8 has not been examined thus far, even though Fe ions,50 zero-valent Fe51 and cobalt ions52 have been broadly studied for such processes. Fe–Co–S based coat exhibits significantly higher efficiency for Fenton-like degradation of MB compared to the FeS2 and CoS2 coat. The mechanism of synergistic effect between different metals in bimetallic systems and their complexes which leads to enhanced heterogeneous Fenton-like performance has been broadly studied for bimetallic oxides53 and so far attracted less attention for bimetallic sulfides.54 The enhancement of degradation activity of synergistically acting bimetallic system is attributable especially to: (i) significant increasing of the active sites on the surface, (ii) accelerating the transfer of electrons and the redox cycle of metal ions which occurs due to the interaction between the oxidation–reduction pairs of iron and cobalt. This redox process can temporarily slow the radical formation and MB decomposition in the initial stage (observed in the first 120 minutes – Fig. 8a violet curve). Consequently, partial surface oxidation of sulfides to sulfates occurs simultaneously with the catalysis of primary radicals, which results in gradual production of sulfate ions and in acceleration of the reaction in later stage.67 Thus well-designed bimetallic oxide and/or sulfide system and their complexes are favourable for forming ˙OH from H2O2 in a wide pH range and provide higher Fenton-like efficiency for organics degradation and also using of H2O2 with limited metal ion leaching and prolonged stability. However, in our case also the contribution of presence of the Fenton-active Fe2O3 (ref. 55) and/or amorphous phase and its potentially synergistic interactions with dominant FeCo8S8 phase may have an additional effect.
 |
| Fig. 8 Comparison of MB depletion under the conditions of Fenton like degradation (a) and day light driven degradation (b) in the presence of Fe–Co–S, CoS2, FeS2 deposits. | |
3.5.2 Photocatalytic MB degradation under day light. The photocatalytic activity of the same 3 coats have been tested under natural day light. The conditions of the photo-catalytic test corresponded to normal, less intensive sunlight (average intensity of sunlight: Φ ∼ 700 lx). Despite such adverse conditions for the photocatalytic process (low sunlight intensity and low surface area of the active area compared to the MB solution volume), the degradation of MB progressed (Fig. 8b). After approximately 2.5 days (3700 minutes), the relative concentration of MB decreased to 82% for FeS2, 40% for CoS2 and 30% for Fe–Co–S deposit (UV-Vis spectra of photocatalytic MB decrease are available in supplemental data Fig. S9–S11†). The photocatalytic performances of CoS2 (ref. 56) and FeS2 (ref. 28) have been broadly studied. Bimetallic cobalt–iron sulfide is considered to possess superior photocatalytic efficiency compared to monometallic CoS2,19 which agrees with the results obtained under very mild conditions. In accordance with previous studies on bimetallic sulfide photo-catalysts the key specificities of FeCo8S8 presented in Fe–Co–S deposits that lead to an increase in photocatalytic activity are assignable to: faster photo-generated electron transfer and also more convenient trajectories of electron and hole transfer57 accompanied by the bimetallic systems tendency to exhibits reducing of band gap value and enhancing visible light response (compared to its monometallic counterparts).57 Additionally, the synergistic effect between nanostructured FeCo8S8 and Fe2O3 should be taken into account, as strongly enhanced photocatalytic performance has been described for the related system of FeS2/Fe2O3 composite.58
3.5.3 Solar-light driven photocatalytic activity for degradation of selected antibiotics. The Fe–Co–S deposit on Ta, which demonstrated superior photocatalytic efficiency for MB degradation compared to FeS2 and CoS2, has been evaluated for its solar light-driven photocatalytic efficiency in the degradation of sulfamethoxazole (SMX) and trimethoprim (TMP). These compounds were selected as representative antibiotics whose residues are frequently detected in wastewater.59–61 The conditions remained the same as for the MB photocatalytic degradation described above (see Section 2.3) but the intensity of the simulated solar light used here was higher (100 klx). Fig. 9 show progress of SMX and TMP depletion respectively. Both antibiotics slowly degrade also under the influence of solar light irradiation when slightly higher degradation degree after 29 h of sun light exposure has been measured for SMX (relative concentration: 65%) than for TMP (relative concentration: 64%). For both antibiotics the presence of the Fe–Co–S deposit enhanced the degree of degradation. After 29 h of solar light exposure, SMX exhibited relative concentration of 61% and TMP of 47%. A detailed comparative study of SMX and TMP degradation under various AOPs, together with their toxicity and degradation by-products, has been described in the literature.59–61 Based on DFT calculations and previous experimental studies, SMX exhibits a higher photodegradation rate than TMP.59–61 Therefore, it is notable that in the presence of the Fe–Co–S deposit, TMP undergoes significantly higher degradation compared to SMX. To compare photocatalytic efficiency of mixed Fe–Co–S with its individual counterparts the same test was performed also with CoS2 and FeS2. A comparison of the influence of different deposits can be clearly seen in Fig. 9 (time-dependent degradation curves are provided in Fig. S12 and S13†). Efficiency of Fe–Co–S for degradation of SMX is comparable with FeS2 (61% and 59% of respectively), while the CoS2 deposit showed the highest degradation level (50% of residual concentration). In case of TMP effect of Fe–Co–S is almost the same as for CoS2 coat (∼47%), whereas significantly higher photocatalytic contributions shows FeS2 (22% of residual TMP concentration).
 |
| Fig. 9 Solar light driven degradation of sulfamethoxazole (a) and trimethoprim (b) showing relative residual concentration after 29 hours of sunlight irradiation in absence of deposits (pure sunlight effect) and in the presence of Fe–Co–S, CoS2, FeS2 deposits on Ta. | |
In contrary to Fenton-like and photocatalytic degradation of MB the Fe–Co–S deposit does not exhibit enhanced effect for SMX and TMP.
A detailed explanation of the differential photodegradation effect towards specific antibiotics will require a separate study. It will be necessary to investigate the contribution of individual phases and surface properties (e.g. surface charge) of the deposits in relation not only to the given antibiotics, but also to the degradation products.
3.5.4 Electrochemical properties. The OER catalytic performance of FeS2,26,27 CoS2 (ref. 29) and Fe–Co–S coats on Ta were evaluated in a typical three-electrode system with an Ar-saturated 1.0 M KOH solution as the electrolyte. The results of the electrochemical studies on all prepared metal sulfides films formed on Ta sheets are presented in Fig. 10. The peak appearing at 1.37 versus reference hydrogen electrode (RHE) in Fig. 10a and b refer to the conversion of Co3+ to Co4+ attributed to the oxidation of CoOOH to CoO2.62 When scanning to potentials higher than 1.5 V versus RHE, the current density rapidly increases due to the OER. The overpotential (η) at a current density of 1 mA cm−2 was employed to evaluate the electrocatalytic activity for the OER. It can be seen that Fe–Co–S deposited interface show a small overpotential of 260 mV (Fig. 10a) as compared to the individual CoS2 deposit. It should be noted that the FeS2 deposit and pure Ta sheets exhibited little or no electrocatalytic activity toward the OER within the potential range presented in Fig. 10a. The Fe–Co–S deposited interface exhibits the lowest Tafel slope of 48 mV dec−1 in the selected potential range (Fig. 10c) and hence required lower energy to activate water as compared to monometallic CoS2 interfaces. Further, the charge transfer resistance of Fe–Co–S deposits was relatively smaller than CoS2 (0.8 ohm – Fig. 10d), indicative of the higher electronic conductivity. Taking into fact that FeS2 exhibit no significant water oxidation activity, it is plausible to state that the higher-valence cobalt species in the Fe–Co–S deposit coupled with higher electronic conductivity activates water at potential above 1.45 V leading to the faster kinetics towards water oxidation and subsequent evolution of oxygen.
 |
| Fig. 10 (a) CV curves with a scan rate of 10 mVs−1 in 1.0 M KOH solution for the pristine Ta sheet, CoS2/Ta, FeS2/Ta, Fe–Co–S/Ta, respectively. (b) Magnified view of CoS2/Ta, FeS2/Ta, Fe–Co–S/Ta voltammetric traces presented in (a). (c) Tafel plots for CoS2/Ta and Fe–Co–S/Ta. (d) Nyquist plots of Ta sheet, CoS2/Ta and Fe–Co–S/Ta samples measured in the frequency range from 100 kHz to 0.1 Hz with an amplitude of 5 mV (peak-to-peak) at open circuit potential. (Inset: Equivalent circuit used for fitting electrochemical impedance spectra; the fitted EIS plot and the raw data points used for fitting are presented together in Fig. S14.†) | |
4. Conclusions
Despite the many attractive properties of bimetallic iron–cobalt sulfides, which exceed the characteristics of monometallic iron and cobalt sulfides and thus expand the application potential, a facile one-step process for the formation of this desirable phase remains challenging. In this contribution, we report our finding that pulsed laser irradiation of equimolar mixed FeS2 and CoS2 targets results in the deposition of nanostructured thin films on Ta substrates. The prepared nanostructured deposits were analyzed by scanning electron microscopy, Raman spectroscopy, X-ray photoelectron spectroscopy, X-ray diffraction, high resolution electron microscopy, and electron diffraction. These complementary analyses revealed that the deposit on Ta consisted of nanocrystalline cobaltpentlandite [FeCo8S8] along the maghemite [γ-Fe2O3] covered with only a superficial layer of cobalt sulfate. The mechanism of FeCo8S8 formation is attributed to the intermixing events of gaseous excited Fe, Co, and S species in laser-induced plasma, followed by the higher deposition tendency of CoS2 in the presence of iron ions/atoms, leading to the creation of the FeCo8S8 phase.
The FeCo8S8-based deposit was tested for its ability to degrade methylene blue under the conditions of daylight photocatalysis and Fenton-like reactions, and for its ability to degrade sulfamethoxazole and trimethoprim under solar light-driven photocatalysis. Catalytic testing showed the superior performance of FeCo8S8-based deposits for MB degradation compared to individual FeS2 and CoS2, whereas the Fenton-like activity of bimetallic FeCo8S8 is revealed here for the first time. Solar light-driven activity of FeCo8S8-based deposit for degradation of SMX and TMP was not increased compared to individual monosulfides. Detailed explanation requires further separate catalytic study. In addition, the electrochemical properties of the FeCo8S8-based deposit toward oxygen evolution reaction (OER) were studied and FeCo8S8 exhibited a significantly higher electrocatalytic alkaline water oxidation activity than FeS2 and CoS2. DFT calculations involving the investigations on thermodynamic stability and electronic structure of mixed FeS2 and CoS2 supported the exclusive formation of cobaltpentlandite [FeCo8S8] along with the maghemite [γ-Fe2O3] phase upon pulse laser irradiation.
Data availability
The source data are available in Zenodo repository with the identifier: https://doi.org/10.5281/zenodo.13853215. Preprint copy available at: http://dx.doi.org/10.2139/ssrn.5169823.
Author contributions
Tomáš Křenek: conceptualization, data curation, investigation, methodology, visualization, writing – original draft, supervision. Lukáš Vála: conceptualization, writing – original draft, visualization, investigation, data curation, supervision. Palaniappan Subramanian: conceptualization, writing – original draft, visualization, investigation, data curation. Saleem Ayaz Khan: conceptualization, writing – original draft, visualization, investigation, data curation. Ján Minár: data curation, investigation, methodology, supervision. Martin Koštejn: data curation, investigation, formal analysis. Rostislav Medlín: data curation, investigation, formal analysis. Petr Mikysek: data curation, investigation, formal analysis. Věra Jandová: data curation, investigation, formal analysis. Veronika Vavruňková: data curation, investigation, formal analysis.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This publication was supported by the project “Mechanical Engineering of Biological and Bio-inspired Systems”, funded as project no. CZ.02.01.01/00/22_008/0004634 by Programme Johannes Amos Commenius, call Excellent Research. The study was supported also by the project no. SGS-2024-022 “Progressive materials and technologies for green transformation” (programme of Ministry of Education, Youth and Sports of Czech Republic); and the study was conducted also under institutional support from the Institute of Geology of the Czech Academy of Sciences (RVO67985831).
References
- S. Chandrasekaran, L. Yao, L. Deng, C. Bowen, Y. Zhang, S. Chen, Z. Lin, F. Peng and P. Zhang, Chem. Soc. Rev., 2019, 48, 4178–4280 RSC.
- M. Chhowalla, H. S. Shin, G. Eda, L. J. Li, K. P. Loh and H. Zhang, Nat. Chem., 2013, 5, 263–275 CrossRef PubMed.
- Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman and M. S. Strano, Nat. Nanotechnol., 2012, 7, 699–712 CrossRef CAS PubMed.
- Y. Wang and N. Herron, J. Phys. Chem., 1991, 95, 525–532 CrossRef CAS.
- D. Voiry, R. Fullon, J. Yang, C. de Carvalho Castro e Silva, R. Kappera, I. Bozkurt, D. Kaplan, M. J. Lagos, P. E. Batson, G. Gupta, A. D. Mohite, L. Dong, D. Er, V. B. Shenoy, T. Asefa and M. Chhowalla, Nat. Mater., 2016, 15, 1003–1009 CrossRef CAS PubMed.
- Y. Liu, Q. Li, R. Si, G.-D. Li, W. Li, D.-P. Liu, D. Wang, L. Sun, Y. Zhang and X. Zou, Adv. Mater., 2017, 29, 1606200 CrossRef PubMed.
- Y. Zhang, T. T. Zuo, J. Popovic, K. Lim, Y. X. Yin, J. Maier and Y. G. Guo, Mater. Today, 2020, 33, 56–74 CrossRef CAS.
- S. Fang, D. Bresser and S. Passerini, Adv. Energy Mater., 2020, 10, 21 Search PubMed.
- P. Z. Chen, Y. Tong, C. Z. Wu and Y. Xie, Acc. Chem. Res., 2018, 51, 2857–2866 CrossRef CAS PubMed.
- S. S. Ren, X. D. Duan, S. Liang, M. D. Zhang and H. G. Zheng, J. Mater. Chem. A, 2020, 8, 6144–6182 RSC.
- R. Rojaee and R. Shahbazian-Yassar, ACS Nano, 2020, 14, 2628–2658 CrossRef CAS PubMed.
- Z. F. Huang, J. Wang, Y. C. Peng, C. Y. Jung, A. Fisher and X. Wang, Adv. Energy Mater., 2017, 7, 21 Search PubMed.
- X. M. Lin, J. H. Chen, J. J. Fan, Y. Ma, P. Radjenovic, Q. C. Xu, L. Huang, S. Passerini, Z. Q. Tian and J. F. Li, Adv. Energy Mater., 2019, 9, 10 Search PubMed.
- S. Lu, J. Jiang, H. Yang, Y. J. Zhang, D. N. Pei, J. J. Chen and Y. Yu, ACS Nano, 2020, 14, 10438–10451 CrossRef CAS PubMed.
- A. Sivanantham, P. Ganesan and S. Shanmugam, Adv. Funct. Mater., 2016, 26, 4661 CrossRef CAS.
- S. Czioska, J. Wang, X. Teng and Z. Chen, ACS Sustainable Chem. Eng., 2018, 6, 11877–11883 CrossRef CAS.
- J. Chen, D. H. C. Chua and P. S. Lee, Small Methods, 2020, 4, 1900648 CrossRef CAS.
- Y. Chen, Z. C. Lai, X. Zhang, Z. X. Fan, Q. Y. He, C. L. Tan and H. Zhang, Nat. Rev. Chem., 2020, 14, 269 CrossRef PubMed.
- G. Gurumoorthy, P. J. Rani, S. Thirumaran and S. Ciattini, Inorg. Chim. Acta, 2017, 455, 132–139 CrossRef CAS.
- Y. Zhou, M. Luo, Z. Zhang, W. Li, X. Shen, W. Xia, M. Zhou and X. Zeng, Appl. Surf. Sci., 2018, 448, 9–15 CrossRef CAS.
- S. Lu, J. Jiang, H. Yang, Y. J. Zhang, D. N. Pei, J. J. Chen and Y. Yu, ACS Nano, 2020, 14, 10438–10451 CrossRef CAS PubMed.
- A. Naderi, X. Yong, M. Karamad, J. Cai, Y. Zang, I. Gates, S. Siahrostami and G. Wang, Appl. Surf. Sci., 2021, 542, 148681 CrossRef CAS.
- X. Hu, S. Liu, Y. Chen, J. Jiang, Y. Sun, L. Wang, S. Han and H. Lin, Ionics, 2020, 26, 2095–2106 CrossRef CAS.
- X. Hu, R. Wang, P. Sun, Z. Xiang and X. Wang, ACS Sustainable Chem. Eng., 2019, 7, 19426–19433 CrossRef CAS.
- Z. Liu, Z. Zhang, Z. Li, X. Xie, S. Zhong, Y. Li, Z. Xu and H. Liu, Chem. Eng. J., 2021, 418, 129442 CrossRef CAS.
- J. Pola, M. Urbanova, D. Pokorná, P. Bezdička, J. Kupčík and T. Křenek, RSC Adv., 2014, 11543–11551 RSC.
- T. Křenek, R. Medlín, S. Karatodorov, V. Mihailov, M. Pola and A. H. Reshak, J. Alloys Compd., 2017, 723, 689–697 CrossRef.
- T. Křenek, L. Vála, T. Kovářík, R. Medlín, R. Fajgar, J. Pola, V. Jandová, V. Vavruňková, M. Pola and M. Koštejn, Dalton Trans., 2020, 49, 13262 RSC.
- L. Vála, R. Medlín, M. Koštejn, S. Karatodorov, V. Jandová, V. Vavruňková and T. Křenek, Eur. J. Inorg. Chem., 2019, 1220–1227 CrossRef.
- J. L. Labar, Ultramicroscopy, 2005, 103, 237–249 CrossRef CAS PubMed.
- A. Vaitkus, A. Merkys, T. Sander, M. Quirós, P. A. Thiessen, E. E. Bolton and S. Gražulis, J. Cheminf., 2003, 15, 1–15 Search PubMed.
- G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS.
- G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
- G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
- D. Ma, B. Hu, W. Wu, X. Liu, J. Zai, C. Shu, T. T. Tsega, L. Chen, X. Qian and T. L. Liu, Nat. Commun., 2019, 10, 3367 CrossRef PubMed.
- https://rruff.info/.
- T. Křenek, T. Stich, D. Docheva, P. Mikysek, M. Koštejn, V. Jandová, R. Fajgar, M. Pola, T. Kovářík, D. Moskal, J. Forejtová and J. Pola, Surf. Coat. Technol., 2020, 447, 128820 CrossRef.
- M. Fantauzzi, B. Elsener, D. Atzei, A. Rigoldi and A. Rossi, RSC Adv., 2015, 5(93), 75953–75963 RSC.
- M. C. Biesinger, B. P. Payne, A. P. Grosvenor, L. W. M. Lau, A. R. Gerson and R. S. C. Smart, Appl. Surf. Sci., 2011, 57(7), 2717–2730 CrossRef.
- I. G. Dance, K. J. Fisher and G. D. Willett, J. Chem. Soc., Dalton Trans., 1997, 2557–2562 RSC.
- F. Aubriet, C. Poleunis, J.-F. Miller and P. Bertrand, J. Mass Spectrom., 2006, 41, 527–542 CrossRef CAS PubMed.
- F. Aubriet and J.-F. Miller, J. Am. Soc. Mass Spectrom., 2008, 19, 488–501 CrossRef CAS PubMed.
- A. V. Bulgakov, A. B. Evtushenko, Y. G. Shukhov, I. Ozerov and W. Marine, AIP Conf. Proc., 2010, 1278, 78–89 CrossRef CAS.
- J. S. McIndoe, Transition Met. Chem., 2003, 28, 122–131 CrossRef CAS.
- S. Karatodorov, V. Mihailov, T. Křenek and M. Grozeva, J. Phys.: Conf. Ser., 2016, 700, 012001 CrossRef.
- H. Ning, Z. Liu, Y. Xie and H. Huang, J. Electrochem. Soc., 2018, 165(9), A1725–A1733 CrossRef CAS.
- A. Teruya, F. Suzuki, D. Aoki, F. Honda, A. Nakamura, M. Nakashima, Y. Amako, H. Harima, K. Uchima, M. Hedo, T. Nakama and Y. Ōnuki, J. Phys.: Conf. Ser., 2007, 807, 012001 CrossRef.
- X. Guo, J. Jia, Y. Xu, Q. Meng, F. Zha, X. Tang and H. Tian, Appl. Surf. Sci., 2021, 556, 149786 CrossRef CAS.
- P. J. Masset and R. A. Guidotti, J. Power Sources, 2008, 178, 456–466 CrossRef CAS.
- D. Wu, Y. Chen, Y. Zhang, Y. Feng and K. Shih, Sep. Purif. Technol., 2015, 154, 60–67 CrossRef CAS.
- N. Wang, T. Zheng, G. Zhang and P. Wang, J. Environ. Chem. Eng., 2016, 4, 762–787 CrossRef CAS.
- W. Fua, J. Yia, M. Cheng, Y. Liu, G. Zhang, L. Lia, L. Dua, B. Lia, G. Wanga and X. Yanga, J. Hazard. Mater., 2022, 424, 127419 CrossRef PubMed.
- B. Ren, Z. Xinga, N. Zhanga, T. Chenga, X. Liua, W. Chena, Z. Wanga, Z. Li and W. Zhou, J. Colloid Interface Sci., 2025, 677, 882–895 CrossRef CAS PubMed.
- S. Zha, Y. Cheng, Y. Gao, Z. Chen, M. Megharaj and R. Naidu, Chem. Eng. J., 2014, 255, 141–148 CrossRef CAS.
- A. D. Bokare and W. Choi, J. Hazard. Mater., 2014, 275, 121–135 CrossRef CAS PubMed.
- L. Li, J. Xu, J. Ma, Z. Liu and Y. Li, J. Colloid Interface Sci., 2019, 552, 17–26 CrossRef CAS PubMed.
- M. M. Mohamed, E. E. Abdelmonem and G. O. El-Sayed, Journal of Water Process Engineering, 2023, 53, 103633 CrossRef.
- S. Oros-Ruiz, et al., J. Hazard. Mater., 2013, 263P, 28–35 CrossRef PubMed.
- S. Oros-Ruiz, R. Zanella and B. Prado, J. Photochem. Photobiol., A, 2018, 360, 95–108 CrossRef.
- O. Mertah, A. Gómez-Avilés, A. Slassi, A. Kherbeche, C. Belver and J. Bedia, Catal. Commun., 2023, 175, 106611 CrossRef CAS.
- P. Sivakumar, P. Subramanian, T. Maiyalagan, A. Gedanken and A. Schechter, Mater. Chem. Phys., 2019, 229, 190–196 CrossRef CAS.
- O. Knop, C. H. Huang, K. I. G. Reid, J. S. Carlow and F. W. D. Woodhams, Chalkogenides of the Transition Elements. X. X-Ray, Neutron, Mossbauer, and Magnetic Studies of Pentlandite and the π Phases π(Fe, Co, Ni, S), Co8MS8, and Fe4Ni4MS8 (M = Ru, Rh, Pd), J. Solid State Chem., 1976, 16, 97–116 CrossRef CAS.
- International Centre for Diffraction Data (ICDD), PDF-4/Axiom Database, Newtown Square, PA, USA, 2023 Search PubMed.
- R. Schrader and G. Büttner, Eine neue Eisen(III)-oxidphase: ε-Fe2O3, Z. Anorg. Allg. Chem., 1963, 320, 220–234 CrossRef CAS.
- L. Díaz, M. Santos and J. Pola, J. Organomet. Chem., 2007, 692, 3841–3845 CrossRef.
- M. Xing, W. Xu, C. Dong, Y. Bai, J. Zeng, Y. Zhou and Y. Yin, Metal Sulfides as Excellent Co-catalysts for H2O2 Decomposition in Advanced Oxidation Processes, Chem, 2018, 4(6), 1359–1372 CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.