Daniel
Garstenauer
ab,
Patrick
Guggenberger
ab,
Ondřej
Zobač
c,
Franz
Jirsa
de and
Klaus W.
Richter
*af
aDepartment of Functional Materials & Catalysis, University of Vienna, Josef-Holaubek-Platz 2, 1090 Vienna, Austria. E-mail: klaus.richter@univie.ac.at
bVienna Doctoral School in Chemistry, University of Vienna, Währinger Straße 42, 1090 Vienna, Austria
cInstitute of Physics of Materials, Czech Academy of Sciences, Žižkova 22, 61600 Brno, Czech Republic
dDepartment of Inorganic Chemistry, University of Vienna, Josef-Holaubek-Platz 2, 1090 Vienna, Austria
eDepartment of Zoology, University of Johannesburg, Auckland Park, 2006 Johannesburg, South Africa
fX-ray Structure Analysis Centre, University of Vienna, Währinger Straße 42, 1090 Vienna, Austria
First published on 9th October 2024
The development and design of catalysts have become a major pillar of latest research efforts to make sustainable forms of energy generation accessible. The production of green hydrogen by electrocatalytic water splitting is dealt as one of the most promising ways to enable decarbonization. To make the hydrogen evolution reaction through electrocatalytic water splitting usable on a large scale, the development of highly-active catalysts with long-term stability and simple producibility is required. Recently, nickel tellurides were found to be an interesting alternative to noble-metal materials. Previous publications dealt with individual nickel telluride species of certain compositions due to the lack of broadly applicable synthesis strategies. For the first time, in this work the preparation of carbon black supported nickel telluride nanoparticles and their catalytic performance for the electrocatalytic hydrogen evolution reaction in alkaline media is presented. The facile vapour–solid synthesis strategy enabled remarkable control over the crystal structure and composition, demonstrating interesting opportunities of active site engineering. Both single- and multi-phase samples containing the Ni–Te compounds Ni3Te2, NiTe, NiTe2−x & NiTe2 were prepared. Onset potentials and overpotentials of −0.145 V vs. RHE and 315 mV at 10 mA cm−2 respectively were achieved. Furthermore, it was found that the mass activity was dependent on the structure and composition of the nickel tellurides following the particular order: Ni3Te2 > NiTe > NiTe2−x > NiTe2.
Hydrogen is expected to be one of the most promising and potent energy carriers to tackle the challenges of climate crisis and substitute fossil fuels.3,4 Hydrogen offers versatile options of implementation in sustainable future scenarios.5 A high gravimetric energy density of 120 MJ kg−1 and eminent abundance make hydrogen suitable for efficient energy storage.6 Among the most promising technologies for clean hydrogen production, electro-catalytic water splitting powered by solar, or wind power has been the centre of attention for decades.7–9 Water electrolysis involves the two half-cell reactions: hydrogen evolution (HER) on the cathode side, with simultaneous oxygen evolution (OER) on the anode side (Fig. 1).
Fig. 1 Scheme of water electrolyser with respective half-cell reactions in acidic, alkaline, and neutral media. |
Unfortunately, both reactions, but in particular OER, are hindered by sluggish reaction kinetics, demanding for high overpotentials to achieve reasonable current densities. Thus, input potentials of practical water electrolysers are much larger than the theoretical 1.23 V, ultimately leading to low energy conversion efficiencies.10 To perform water splitting with acceptable efficiency, sophisticated electrocatalysts are needed. State-of-the-Art catalysts involve noble metal-based materials (e.g. Pt, IrO2 and RuO2) for acidic media and transition metal-based materials (e.g. Ni, Fe and Co) for alkaline media.9,11 Though, high cost and unsatisfactory catalytic stability of noble metal-based catalysts as well as moderate activity of transition metal-based catalysts yet hinder the broad application of water splitting for industrial-scale H2 production. Hence, realization of large-scale H2 production by water splitting requires the development of competitive non-noble electrocatalysts. On account of favoured thermodynamics for OER in alkaline media, potential catalysts for HER should also be optimised for alkaline media to allow reconcilable implementation.
Recently, “active site engineering”, in which the accessibility, reactivity and density of catalytically active sites should be controlled through optimized synthesis approaches, became the significant strategy for catalyst development.12 Latest research focusses on the idea of morphology and electronic structure tuning.13,14 Morphology tuning strategies target the formation of 2D and highly dispersed 3D materials (e.g. nanoparticles). Those materials have higher specific surface areas and porosities, therefore an increase in accessibility of active sites compared to planar bulk materials is achieved. Electronic structure tuning on the other hand utilises the possibility to alter the electronic structure and form active sites through vacancy engineering, elemental doping, or control over crystal phase formation.
Following these developments, intermetallic compounds moved into the spotlight of researchers.15 Intermetallic compounds consist of two or more (transition-)metals arranged in crystal structures differing from their constituting metals (e.g. transition metal chalcogenides). In the intermetallic compounds a mixture of metallic, covalent, and ionic bond contents leads to unique physical, chemical, and mechanical characteristics.16 Long-range order, high stability and designability of morphology and electronic structure make intermetallic compounds ideal catalyst materials.17 Although their excelling catalytic performance is well documented, synthesis of intermetallic compounds, particularly in nanoparticle shape remains challenging.18
Synthesis methods of intermetallic nanoparticles can typically be divided into the two categories of thermal annealing and wet-chemistry approaches.19 For thermal annealing routes, thermodynamic and kinetic studies imply that disordered alloys exist in a meta-stable state relative to the well-ordered intermetallic phase. When the alloy is annealed at elevated temperatures the atoms will arrange in the highly ordered intermetallic structure.20 Unfortunately, the high temperatures needed to overcome the kinetic energy barriers promote particle aggregation, sintering and Ostwald ripening. In wet-chemistry methods, on the other hand, co-reduction of dissolved precursors is carried out in the presence of stabilizers, capping- and reduction agents at milder temperatures. The need for various additives not only limits the product quality but also enforces difficult individual synthesis development for each intermetallic system, limiting their comparability and hindering their broad usage.21
A preeminent response to the above-mentioned synthesis challenges is the facile, direct vapour–solid (VS) synthesis method. The vapour–solid synthesis emerged from the isopiestic method and was originally developed for the stoichiometric preparation and study of intermetallic bulk compounds.22,23 Subsequently, supported intermetallic nanoparticles moved into the focus of research effort in our group.24,25 For the preparation of intermetallic compounds by the vapour–solid approach, stoichiometric amounts of the constituting metals are sealed in an evacuated quartz vessel, separated in space and equilibrated within an temperature gradient. The reaction partner with lower vapour pressure condenses on the coolest part of the system. From the respective temperature the system pressure and consequently the activity are controlled (Fig. 2). The absence of any additives, use of simple, commercially available educts with high purity and the excellent compositional control enables the synthesis of various intermetallic compounds with morphological and electronic structure tunability.
Nickel-based materials can offer highly active and stable alternatives to state-of-the-art catalysts.26 A comparison between representative nickel chalcogenide materials and their reported overpotential at 10 mA cm−2 are presented in Table 1. In particular Nickel-tellurides unveiled outstanding electrocatalytic performance for HER in recent publications.27–41
Nonetheless, the research field lacks a broadly applicable synthesis strategy capable of distinct and composition-controlled preparation. To the best of our knowledge, this is the first time the synthesis of carbon black supported nickel telluride nanoparticles and their performance in electrocatalytic HER are reported. In this work we aim to present the direct vapour–solid method to a wide audience and highlight its potential in future catalyst development by benchmarking carbon black supported, single- and multiphase nickel telluride nanoparticles for electrocatalytic HER in alkaline medium. An overview about the intermetallic nickel–tellurium system43 from bulk studies is presented in Table 2.
Name | Formula | Structure type | Space group | Homogeneity range43 [at% Te] | Lattice parameters [nm] | References |
---|---|---|---|---|---|---|
(Ni) | Ni | Cu | Fmm | — | a = 3.5181 | P. Juhás et al.44 |
β2 | Ni3+xTe2 | Own type | P121/m1 | 41–42 | a = 0.3376 | J. Barstard et al.45 |
b = 0.3800 | ||||||
c = 0.6100 | ||||||
β = 91.25° | ||||||
γ1 | NiTe0.775 | Own type | Pmc21 | — | a = 0.3916 | K. Klepp et al.46 |
b = 0.6860 | ||||||
c = 1.232 | ||||||
δ | NiTe2−x | NiAs | P63/mmc | 51–67 | a = 0.3956 | J. Barstard et al.45 |
c = 0.5370 |
All chemicals were used as purchased. Water was purified and deionized (18.2 MΩ cm) in-house by a Milli-Q® (Merck, Germany) system.
Phase composition and structural information of the catalysts was acquired by powder X-ray diffraction with a D8 Advance (Bruker, Germany) diffractometer in Bragg–Brentano pseudo focusing mode with theta/theta geometry and structure refinement with the Topas 7.13 software. Zero-background silicon single-crystal plates covered with a thin layer of grease to fix the powders were used as sample holders. For all measurements, the diffractometer was mounted with a one-dimensional silicon strip detector, Lynxeye (Bruker, Germany). Patterns were determined in the 30° to 90° 2-theta range over 3 h with an acceleration voltage of 40 kV and a beam current of 40 mA.
Nitrogen adsorption–desorption isotherms were recorded at −196 °C with a QuantaTech Inc. iQ3 (Anton Paar, USA) physisorption setup. Outgassing was performed at 80 °C for 12 h. The Anton Paar QuantaTech Inc. ASiQWin 5.2 software was used for data evaluation. The Brunauer–Emmett–Teller (BET) equation was applied to calculate the specific surface area (SSA) in the relative pressure range of 0.05–0.20 P/P0. For comparison, the specific surface area and the pore size distribution were also determined according to the quenched solid density functional theory (QSDFT) from the adsorption branch of the isotherms, which is reported to result in more precise values for carbon materials.48–50 For the QSDFT calculations, slit/cylindrical pore geometries of carbon were assumed. Based on the Gurvich rule, the total pore volume was evaluated at the plateau region of the isotherm at P/P0 = 0.95.
Electron microscopy was carried out for size verification of the intermetallic nanoparticles as well as elemental mapping. The scanning and high-resolution images were taken at a Talos F200i (Thermo Scientific™, Germany) transmission electron microscope (TEM) operated at a voltage of 200 kV, equipped with a field electron gun (FEG) and a 4k·4k Ceta 16M camera. Data evaluation was done in Velox software.
The catalyst suspension was prepared by dispersing 4.8 mg of the catalyst powder in a mixture of 700 μL ultra-pure water, 250 μL 2-propanol, and 50 μL Nafion® solution and subsequent sonication for 15 minutes. The GC-RDE was polished over Polish cloth with 1 μm and 0.05 μm Al2O3 suspension to give a mirror finish and subsequently sonicated in ultra-pure water for 5 minutes. 10 μl of the freshly prepared catalyst suspension were drop-cast onto the polished electrode disc and dried under light irradiation for 30 minutes. The preparation led to a final catalyst loading of the working electrode of 0.24 mg cm−2.
All electrochemical measurements were recorded at a RDE rotation speed of 1600 rpm. For the cyclic voltammetry (CV) and linear sweep voltammetry (LSV) measurements, 90% of the iR-drop was corrected by the Nova 2.1.5 software (Metrohm, Germany). To ensure reproducibility, the measurements were repeated in triplicates. For data evaluation the measurement closest to the median values was chosen. Prior to the catalyst screening, the open circuit potential (OCP) was screened for at least six minutes, the electrode was accepted as stable if potential changes of ≤10 mV s−1 were achieved for the last minute of measurement.
The working electrode was activated by a series of CV measurements under constant argon purging of the electrolyte: alternately, in the potential range of 0.4 to −0.4 V vs. RHE, 1 cycle with a scan rate of 10 mV s−1 was followed by 25 cycles with a scan rate of 100 mV s−1 until 53 activation cycles were recorded in total. Potentiostatic electrochemical impedance spectroscopy (EIS) in a range from 100 kHz to 100 mHz with an amplitude of 5 mV at an applied potential of −0.3 V vs. RHE, allowed for further investigation of the catalyst's reaction kinetics. The electrochemical active surface area (ECSA) was assessed similar as described by McCrory et al.11 by investigating the double layer capacitance: CVs with ascending scan rates (20 to 220 mV s−1, 5 cycles per scan speed) in the non-faradaic current region (0.05–0.15 V vs. RHE) were measured. The respective anodic and cathodic current values at 0.10 V vs. RHE were plotted over the scan rate. The determined double-layer capacitance (CDL) of the system was taken as the average of the absolute values of the slope of the linear fits to the data. The specific capacitance (CS) reference value of 0.040 mF cm−2, which was reported to be a typical approximation in alkaline media, was used to estimate the ECSA as the product of CDL·CS−1.34
To exclude the impact of hydrogen formation during the reaction, the LSV measurements, in the potential range of 0.4 to −0.4 V vs. RHE, were recorded after purging the electrolyte with hydrogen gas for 30 minutes.
Catalyst stability was investigated by chronoamperometric (CA) measurements in argon purged electrolyte over 12 h at a constant potential of −0.25 V vs. RHE. Before and after the CA measurement a LSV curve was recorded and each 500 μL of the electrolyte solution were taken for TXRF spectroscopy investigation.
In addition, TXRF was used to investigate the electrolyte solution used for electrocatalytic tests. Typical limits of detection are in the low ppb range. The spectra did not show any Ni, Te, or Fe impurities prior to catalytic testing, nor after the twelve hours lasting chronoamperometric measurements.
Fig. 3 Measured and refined pXRD patterns with the most significant peaks highlighted of Ni/C (a), Ni39–Te61/C (b) and Ni67–Te33/C (c). |
Sample | ω TXRF(Ni) [wt%] | ω target(Ni) [wt%] | ω TXRF(Te) [wt%] | ω target(Te) [wt%] |
---|---|---|---|---|
Ni85–Te15/C | 17.6 | 15.2 | 6.0 | 5.8 |
Ni78–Te22/C | 19.6 | 19.5 | 8.8 | 11.9 |
Ni67–Te33/C | 9.6 | 6.7 | 9.4 | 7.2 |
Ni60–Te40/C | 8.6 | 6.5 | 10.6 | 9.5 |
Ni49–Te51/C | 15.6 | 11.8 | 28.7 | 26.9 |
Ni39–Te61/C | 12.3 | 10.5 | 39.8 | 35.0 |
Sample | Phase | Lattice parameters [Å] | LVol-FWHM [nm] |
---|---|---|---|
a LVol-FWHM value not observed due to overlapping composition sets of the same phase. | |||
Ni/C | Ni | a = 3.529(9) | 27.(9) |
NiO | a = 4.177(0) | 1.(2) | |
Ni85–Te15/C | Ni | a = 3.528(7) | 45.(1) |
NiO | a = 4.177(0) | 1.(0) | |
NiTe2−x | a = 3.939(2)–3.975(2), c = 5.321(5)–5.368(1) | ||
Ni78–Te22/C | Ni | a = 3.529(3) | 37.(0) |
NiO | a = 4.177(0) | 4.(0) | |
NiTe2−x | a = 3.874(3)–3.963(7), c = 5.329(2)–5.372(0) | ||
Ni67–Te33/C | Ni | a = 3.531(3) | 54.(8) |
NiO | a = 4.177(0) | 1.(3) | |
NiTe2−x | a = 3.977(4), c = 5.368(2) | 36.(1) | |
Ni3Te2 | a = 7.592(2), b = 3.777(1), c = 6.142(5) | 26.(2) | |
Ni60–Te40/C | Ni | a = 3.530(5) | 47.(2) |
NiO | a = 4.177(0) | 1.(2) | |
NiTe2−x | a = 3.948(8)–3.975(5), c = 5.369(2)–5.371(2) | ||
Ni3Te2 | a = 7.586(3), b = 3.770(3), c = 6.136(3) | 22.(9) | |
Ni49–Te51/C | NiO | a = 4.177(0) | 1.(3) |
NiTe2−x | a = 3.902(7)–3.953(0), c = 5.356(1)–5.369(6) | ||
Ni39–Te61/C | NiO | a = 4.177(0) | 1.(2) |
NiTe2−x | a = 3.889(7), c = 5.340(1) | 89.(0) |
Surprisingly, NiTe2−x was present in all samples. Furthermore, it was observed that peak-broadening occurred for those NiTe2−x patterns of samples, which were not targeting the NiTe2−x phase region directly. The peak-broadening is caused by the presence of slightly differing lattice parameter sets reflecting compositional variations, hence hinting that those samples were not finally in thermodynamic equilibrium. Furthermore, the peak broadening hinders an explicit evaluation of the crystal domain size of NiTe2−x in the sample. The NiAs structure type of the NiTe2−x phase has a broad, potential homogeneity range and would allow for intermetallic Ni–Te compositions to extend from Ni2Te to NiTe to NiTe2, dependent on the occupation of vacancies and interstitial sites.57 The lattice parameter ‘a’ is increasing with decreasing tellurium concentration in NiTe2−x.43 For nanoparticles, the phase stability is greatly dependent on morphology and size. A homogeneity range, and the associated flexibility regarding composition and activity, could therefore benefit the stability of such phases compared to stoichiometric line compounds like NiTe0.775. Another factor of this behaviour might be that simpler and comparatively smaller structures are more facile to form in nanoparticles contrary to complex structures with high unit cell volume. Our observations indicate that NiTe2−x is the most favourable structure formed from our starting material.
Rietveld refinement confirmed vacancy formation on the nickel sites in samples with high tellurium concentration, while the nickel doping on interstitial sites could not be seen. This led to a composition of NiTe1.9 and NiTe1.8 for phase-pure catalysts. In samples which were not in equilibrium (e.g. Ni85Te15/C & Ni78Te22/C), NiTe was the dominant NiTe2−x composition. A slight shift towards higher tellurium concentration is caused by partial oxidation of nickel reflected by the observation of almost amorphous NiO in the patterns.
Fig. 4 TEM images of Ni39–Te61/C particle (a) & (b) as well as respective nickel (c) and tellurium (d) elemental maps. |
Energy dispersive X-ray spectroscopy (EDS) confirmed the expected nanoparticle compositions. The nickel nanoparticles of the Ni/C educt were at least superficially oxidized in contact to air, confirming the necessity of pre-reduction and handling under inert atmosphere for the vapour–solid reactions. Selective tellurium intake into the nickel nanoparticles was assured by EDS elemental mapping of the intermetallic Ni–Te/C catalyst powders (Fig. 4).
For particle size analysis, the diameters of about 400 particles per sample were measured in both the horizontal and vertical directions. To minimize the impact of outliers, several powder particles were compared. Representative particle size distributions are shown in Fig. 5. The Ni/C educt displayed a bimodal particle size distribution with estimated diameters of 11 nm and 45 nm, with the smaller particles being the dominant species. The intermetallic Ni–Te/C catalyst powders, on the other hand, showed a more uniform particle size distribution which was shifted slightly in comparison to the educt pattern towards bigger particle sizes. Thus, hinting on the agglomeration and sintering of the small, accumulated nickel particles during the formation of the intermetallic compound. As expected, a higher tellurium concentration leads to the increase of particle size. Very rarely single particles of sizes up to 225 nm were observed in the intermetallic products.
QSDFT calculations performed on the adsorption branch of the isotherm and assuming slit/cylindrical pore geometries of carbon, revealed the presence of micropores in all samples. The average micropore volume, total pore volume and specific surface areas are presented in Table 5. Comparison of the calculated QSDFT and BET specific surface areas lead to comparable values and an expected trend of decreasing surface areas with increasing (inter-)metallic content. The significantly higher surface areas of the samples Ni67–Te33/C and Ni60–Te40/C results from the comparably lower initial metal loading of the respective educt batch.
Sample | S BET [m2 g−1] | S DFT [m2 g−1] | V tot [cm3 g−1] | V micro [cm3 g−1] |
---|---|---|---|---|
Pt/C | 1198 | 992 | 0.84 | 0.25 |
C | 736 | 760 | 0.65 | 0.21 |
Ni/C | 662 | 622 | 0.60 | 0.15 |
Ni85–Te15/C | 483 | 452 | 0.42 | 0.12 |
Ni78–Te22/C | 440 | 402 | 0.40 | 0.10 |
Ni49–Te51/C | 411 | 380 | 0.38 | 0.09 |
Ni39–Te61/C | 415 | 384 | 0.37 | 0.10 |
Ni67–Te33/C | 788 | 746 | 0.72 | 0.20 |
Ni60–Te40/C | 779 | 749 | 0.70 | 0.19 |
Fig. 6 shows the initial and final CV curve of Ni85Te15/C as a representative sample as well as the polarization curves normalized to the geometric surface area (Sgeo), the mass activity and Tafel plots of the prepared catalysts. Besides the onset of the HER at potentials more cathodic than −0.1 V vs. RHE, the CV curves of the nickel-based materials did not show any significant peak in the potential range of interest, implying the absence of redox-reactions of the catalyst material. The CV curves of the other samples are presented in Fig. S3.†
The area normalized polarization curves reveal a positive effect of tellurium addition to a certain extent, but do not take the different nickel loadings of the electrode into account. Hence, the resulting performance is a counterpoise between the positive effect of tellurium addition and the lower absolute nickel concentration in the electrode. To overcome those limitations and get a deeper understanding on the effect tellurium intake has on the catalytic activity, the mass activity is a relevant parameter. Therefore, the polarization curves were additionally normalized to the nickel contents of the catalysts. Comparing the mass activities regarding sample composition reveals that the catalytic performance does not strictly follow the tellurium concentration but the formation of different intermetallic Ni–Te phases (see Table 3). The single-phase samples, Ni39Te61/C and Ni49Te51/C, have the highest tellurium concentrations and consist of the δ phase compound NiTe2−x with low nickel occupation.57 Their polarization curves do not reach the desired value of 10 mA cm−2 in the potential range of interest, but as a trend, the catalytic activity increases with the nickel occupation. For the mixed-phase samples Ni85Te15/C and Ni78Te22/C a positive effect due to the presence of NiTe, a compound of the δ phase with high Ni occupation, was observed. The highest mass activities were achieved due to the presence of Ni3Te2 in the samples Ni67Te33/C and Ni60Te40/C. We conclude that the catalytic activity of the Ni–Te nanoparticles in 1 M KOH follows the particulate order of Ni3Te2 > NiTe > NiTe2−x > NiTe2. This trend appears to agree with literature when comparing different bulk studies, although a comprehensive comparison of all the intermetallic nickel tellurides was not published yet.
It is widely accepted that the HER takes place on the surface of the electrode by a multi-step electrochemical process including:62
H2O + e− + M → M–Hads + OH− (Volmer) | (1) |
H2O + e− + M–Hads → M + H2 + OH− (Heyrovsky) | (2) |
2M–Hads → H2 + 2M (Tafel) | (3) |
Here “M” refers to an active site of the electrocatalyst.
For the understanding of the reaction kinetics, Tafel plots were investigated in the high overpotential region of −0.4 to −0.3 V vs. RHE. The Tafel plots are presented in Fig. 6. The observed Tafel slopes >120 mV dec−1 for all nickel-based catalysts indicate the Volmer step as rate determining step of the HER process. The reaction is therefore limited by the hydrogen adsorption and charge transfer rate of the catalyst. Those observations are supported by the results of electrochemical impedance spectroscopy (EIS) (Fig. 7).
For each nickel-based catalyst only one time constant is observed in the Bode phase angle plot, with comparably high maximum phase angles between 54 to 81° at low peak frequencies in the regime of 1 to 4 Hz. Generally, higher maximum phase angles and lower peak frequencies are correlated to more sluggish reaction kinetics.63 Interestingly, from that observation the Ni/C educt shows the fastest kinetics of the nickel based catalysts. For the intermetallic nickel tellurides again a higher nickel occupation in the δ phase occurs advantageous, NiTe furthermore shows improved HER kinetics compared to Ni3Te2. These observations highlight that the overall catalytic behaviour of the catalysts are not only determined by the reaction kinetics, but by the combination of many factors, like the surface admittance, charge transfer resistance of the interface, active surface area and the accessibility of the active sites.
The admittance acts as a measure of how facile an interface is for electron crossover. It was calculated as the reciprocal of the modulus of impedance by extrapolating the linear region in the Bode absolute impedance plot towards the lowest frequency of the experiments, where the charge transfer is solely controlled by the charge transfer resistance.63
The charge transfer resistance, on the other hand, was obtained by fitting the Nyquist plots with a modified Randel's circuit.51 Bigger diameters of the semi-circles are related to increased charge transfer resistances. The presence of only one semi-circle per sample suggests that the reaction kinetics are dominated by one charge transfer step. Thus, EIS also indicates the charge transfer in the Volmer step to be the rate determining step of the process. As expected, the charge transfer resistance generally follows a contrary trend to the admittance and is increasing with the tellurium intake into Ni/C. Only the NiTe containing Ni85Te15/C sample has a lower charge transfer resistance and higher admittance than Ni/C.
The electrochemical active surface area (ECSA) was estimated from the double layer capacitance through assessment of CV measurements in the non-faradaic region (0.05–0.15 V vs. RHE) at different scan rates (Fig. S4; Table S3†).11 The ECSA is dependent on the material loading as well as the crystal structure, and subsequently on the phase composition of the catalysts. Electrochemical data of the benchmarked catalysts are listed in Table 6. The comparison of the EIS and ECSA results helps to further understand the catalytic performance of the materials. While the admittance follows a contrary trend to the charge transfer resistance, the observed ECSA data does not align with the charge transfer parameter trend. Neither the charge transfer parameters, nor the ECSA data by itself are suited to explain the catalytic activities. The ECSA presents the total accessible surface for electrochemical reactions but does not provide information about the activity of the catalytically active sites. As mentioned above, the Volmer step, which was found to be the rate determining step, is not solely influenced by the charge transfer, but also the hydrogen adsorption activity of the catalytic sites. The charge transfer parameters of NiTe2(−x) samples hint strong oppression, despite the high ECSA ultimately causing the weakest HER performance. The NiTe sample, on the other hand, showcased the most beneficial kinetics and charge transfer parameters, but the respective low ECSA, despite the minimal tellurium intake, hints meagre accessibility of the active sites. Thus, it is concluded that the formation of the intermetallic Ni3Te2 structure significantly improves the accessibility of the active nickel sites for HER in alkaline media. The increased accessibility in combination with good kinetics allow Ni3Te2 to outmatch the other nickel-based catalysts. The catalytic performance of the nickel tellurides follows the order mentioned before, Ni3Te2 > NiTe > NiTe2−x > NiTe2.
Sample | η 10 [mV] | η os [mV] | b [mV dec−1] | ω pf [Hz] | Y [mS] | Θ max [°] | R ct [Ω] | ECSA [m2 g−1] |
---|---|---|---|---|---|---|---|---|
Ni/C | 342 | −155 | 171 | 4.0 | 5.7 | 53.6 | 145 | 46 |
Ni85–Te15/C | 315 | −158 | 150 | 4.0 | 6.3 | 59.2 | 116 | 31 |
Ni78–Te22/C | 331 | −149 | 172 | 2.5 | 5.0 | 56.0 | 186 | 67 |
Ni67–Te33/C | 352 | −145 | 157 | 2.0 | 4.3 | 64.0 | 204 | 60 |
Ni60–Te40/C | 360 | −146 | 144 | 2.5 | 3.4 | 69.1 | 283 | 29 |
Ni49–Te51/C | >400 | −183 | 154 | 1.0 | 2.1 | 69.9 | 752 | 57 |
Ni39–Te61/C | >400 | −229 | 141 | 0.8 | 0.9 | 80.5 | 5302 | 18 |
The absence of Ni, Te, and Fe traces in the TXRF spectra does not only hint on chemical stability of the materials, but also confirms that the “iron effect”,64i.e. the change of activity through the doping with iron from electrolyte impurities, is excludable.
The chronoamperometry investigations depicted in Fig. 8 reveal two zones. An initial rapid decrease of the current density (t < 1 h) is followed by steady state conditions. The rapid catalyst aging effect is known for nickel catalysts,65 possible causes include the charging of the electrical double layer, adsorption/desorption processes, surface restructuring and mass transport changes. The zig-zag form of Ni85Te15/C curve is caused by the formation of H2 gas bubbles which are not removed efficiently from the electrode surface, this joins the observation of fast kinetics of the sample and highlights mass transport limitations in the particular case. An interesting trend is shown by the NiTe2(−x) samples, as Ni39Te61/C and Ni49Te51/C become gradually activated over time. Possibly, surface reconstruction of the telluride species leads to lesser vacancy occupation which was found to be more active towards HER as described before.
Fig. 8 Chronoamperometry measurements at −0.25 V vs. RHE and 1600 rpm RDE rotation rate (a), polarization curves before chronoamperometry (solid lines) and after chronoamperometry (dashed lines) (b). |
The initial rapid aging effect was also reflected in the comparison of the polarization curves before and after the 12 h chronoamperometric run. Besides Ni39Te61/C, for all catalysts a significant increase in the HER overpotentials was found. These observations lead to the conclusion that the catalysts are chemically stable and remain active during the reaction period in alkaline media, but the durability suffers from early deactivation processes.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nr03397c |
This journal is © The Royal Society of Chemistry 2024 |