Mohan Paudela,
Sanjit Karki
a,
Narayan Acharya
a,
Sashil Chapagain
a,
Johann V. Hemmer
a,
Dillon T. Hofsommer
a,
Gautam Gupta
*b,
Robert M. Buchanan
*a and
Craig A. Grapperhaus
*a
aDepartment of Chemistry, University of Louisville, 2320 S. Brook St, Louisville, KY 40292, USA. E-mail: robert.buchanan@louisville.edu; grapperhaus@louisville.edu
bDepartment of Chemical Engineering, University of Louisville, Louisville, Kentucky 40292, USA
First published on 27th February 2025
Green hydrogen, generated through the electrolysis of water using renewable energy sources, is recognized as a highly promising alternative to fossil fuels in the pursuit of net-zero carbon emissions. Electrocatalysts are crucial for reducing overpotentials and enhancing the efficiency of the hydrogen evolution reaction (HER) for the production of green hydrogen. Homogeneous HER serves as a primary method to assess the activity and mechanisms of novel non-precious molecular electrocatalysts in pursuit of replacing precious platinum standards. However, these catalysts can sometimes exhibit instability under reductive and acidic conditions during homogeneous HER. Thus, it is also essential to evaluate catalysts through heterogeneous HER for initial assessment and practical application. In this study, we examine a series of structurally related N2S2 chelated Ni(II) complexes, which are tailored to optimize the basicity of the catalyst for heterogeneous HER activity. These complexes are insoluble in 0.5 M H2SO4, and the films formed after catalyst deposition on glassy carbon electrodes (GCEs) exhibit catalytic currents during HER, demonstrating moderate to good overpotentials, Tafel slopes, and charge transfer resistance. Furthermore, we observe the anticipated structure–activity relationship that arises from tuning the catalyst structure. The complexes maintain stability over extended reductive cycling, as confirmed by various surface characterization techniques, including SEM, EDX, XPS, and XRD. This study highlights the potential of utilizing catalyst basicity to develop efficient and robust heterogeneous HER catalysts.
Considerable effort has focused on the use of transition metal catalysts,7 non-metal carbon-based catalysts,8–10 metal alloys,8,9 metal–organic frameworks,9,11 and metals and mixed metal oxides,12 phosphides9,13 or nitrides14,15 as HER electrocatalysts. Molecular catalysts containing inexpensive earth-abundant metals coordinated by one or more ligands are of particular interest as their physical and electronic structures can be tuned to modulate catalytic activity, which makes them excellent candidates for structure–activity studies.16 Our group and others have studied redox active thiosemicarbazone-based molecular catalysts incorporating earth-abundant metals such as nickel, copper, and zinc as promising homogenous electrocatalysts demonstrating ligand-assisted metal-center,17–19 metal-assisted ligand-centered,20,21 and ligand-centered22,23 HER mechanisms as a function of the metal and the ligand. A significant concern of homogeneous molecular HER catalysts is their potential instability in solution under acidic and/or reducing conditions, which can yield heterogenous, catalytically active degradation products.24–26
Recently, we reported the homogeneous HER activity of six structurally related Ni(II)N2S2 complexes with bis(thiosemicarbazonato) (BTSC) (Ni-1, Ni-4), thiosemicarbazonato-alkylthiocarbamato (TSTC) (Ni-2, Ni-5), and bis(alkylthiocarbamato) (BATC) (Ni-3, Ni-6) ligands (Fig. 1) in acetonitrile using acetic acid as proton donor.27–29 The complexes in that study displayed a systematic decrease in reduction potentials, progressing from NiBTSC to NiTSTC to NiBATC. The computational study indicates over 91% of the spin density of mono reduced derivative [NiL−] of all the complexes is localized on the ligand indicating the first reductions are ligand centered. This trend is accompanied by a reduction in ligand basicity as quantified by pKa measurements using hydrogen triphenylphosphonium tetrafluoroborate as proton source (see Fig. 1). Computational exploration of the heteroatoms demonstrates protonation is most favored at the hydrazino nitrogen resulting in the formation of [NiLH]. Despite these structural and electronic differences, Ni-1–Ni-6 exhibited nearly identical HER overpotentials that were similar to the Ni(II) salt Ni(OTf)2. These results strongly suggested decomposition of the homogeneous catalysts upon exposure to a combination of reducing and acidic conditions.30 In the current study, complexes Ni-1–Ni-6 were immobilized on an electrode surface to evaluate how translation from homogeneous to heterogeneous conditions influenced HER activity and catalyst stability. The immobilized catalysts were evaluated before and after HER using a variety of spectroscopic, microscopic, and electrochemical methods.
The evolution of H2 was confirmed using an H-cell fitted with a gas-tight “low-volume cell cap kit” (Pine Research) equipped with an Ag/AgCl reference, gas dispersion tube, gas outlet, and the working electrode coated with catalyst. The Pt mesh counter electrode was separated from the working electrode compartment by a Nafion 115 membrane, and each side was filled with 0.5 M H2SO4. Chronopotentiometry at −10 mA cm−2 was performed for 120 minutes under a continuous stream of N2 at a flow rate of 15 sccm and 1 mL of gas was sampled every 5 minutes by an in-line gas chromatographer (SRI Instrument, Multiple Gas Analyzer #1 + Sulfur). faradaic efficiencies were quantified by comparing the number of moles of charge used for H2 production with the total number of moles of electrons supplied during the measurement.
Scanning electron microscopy (SEM) images were obtained using an Apreo 2 C field emission scanning electron microscope (Thermo Fischer Scientific). Images of the catalyst materials before and after 1000 reductive CV cycles were obtained by removing the disk tip of the modified electrode. An Everhard-Thornley detector (ETD) in secondary electrode mode was used to analyze the catalyst surface. Samples were analyzed normal to the electrode beam, 2 mm from the cone. Accelerating high voltages and beam current were 2 kV and 6.3 pA, respectively. Elemental analysis and imaging were conducted using an Apreo 2 C field emission scanning electron microscope (Thermo Fischer Scientific) using a QUANTAX Energy-Dispersive X-ray (EDX) detector (Bruker). A voltage of 20 kV and a spot size of 10 were used for imaging, while a primary energy of 40 keV was used for EDX spectrometry. Images were collected 8.5 mm normal from the sample surface.
X-ray photoelectron spectroscopy (XPS) studies were performed using a VG Scientific Multilab 300 equipped with an AlKα (1486.6 eV)/MgKα (1253.6 eV) X-ray twin source, acceleration voltage of 10 kV, and emission current of 12 mA for the X-ray source. The electron energy analyzer operated in Fixed Analyzer Transmission (FAT) mode with a constant energy of 50 eV passed for survey and 20 eV for high-resolution scans.
Modified electrodes 4 and 5 contain molecules with at least one NH-methyl pendent as in 1 and 2, but with the substitution of a methyl group in the ligand backbone with a phenyl group. As shown in Fig. 2a, 4 and 5 show similar improvements in overpotential converging to 0.52 V from initial 0.70 V and 0.76 V, respectively, after peak reductive cycling. In comparison to 1 and 2, the overpotentials for 4 and 5 have increased by ∼0.12 V. Interestingly, 3 and 6, which contain BATC complexes with electron-withdrawing O-ethyl pendent groups instead of electron-donating NH-methyl groups, show no significant improvement in overpotential even after 1000 CV cycles. In addition, 3 and 6 display similar overpotentials (0.670 V at peak reductive cycle) despite different ligand backbone substituents on the immobilized complexes.
The observed changes in overpotential for 1–6 upon reductive cycling are consistent with modification of the ligand structure of the immobilized complexes (Fig. 2b). Notably, electrodes with complexes that are easiest to reduce, 3 and 6, have the highest overpotential. Rather, the overpotential for the optimized electrodes 1–6 depends on the basicity and steric accessibility of the hydrazine nitrogen in the catalyst. Complexes Ni-3 and Ni-6 have non-basic O-Et pendent groups,30 and the largest overpotentials. The four other catalysts all have at least one pendent –NHMe group that is basic with pKa values between 6.88 and 7.61.30 The lowest overpotential is observed with 1 and 2 when the pendent –NHMe group is part of a ligand with only Me groups in the ligand backbone. Overall, the results indicate that peak overpotential is determined by the ease of protonation of the immobilized complex.
Hydrogen production was confirmed through gas chromatography and faradaic efficiencies for 1–6 were calculated using the methodology described in the Experimental section. Overpotentials for each electrode remained relatively constant over 75 minutes and faradaic efficiency values were found to be in the range of 86–92%, which is comparable to values measured for platinum (94%) (Fig. S36–S42†). The kinetics associated with HER for 1–6 were evaluated based on their Tafel slopes. Theoretically, a Tafel slope of 30, 40, or 120 mV dec−1 corresponds to a rate-determining Tafel step (H* + H* → H2), Volmer step (H+ + e− → H*), or Heyrovsky step (H* + H+ + e− → H2), respectively. Deviation from these theoretical values can occur with the surface coverage of the catalyst and the applied potential.34 The Tafel slopes for 1–6 after peak reductive cycling lie between 80 and 130 mV dec−1, Fig. 3a and Table 1. This is most consistent with a Heyrovsky rate-determining step. Data showing the effect of reductive cycling on the Tafel slope are provided in Fig. S9–S14 and Table S2.†
![]() | ||
Fig. 3 (a) Tafel slope and (b) Nyquist plot of 1–6 in comparison to GCE, Pt, and Ni(OTf)2 standards (after peak reductive cycling in 0.5 M aq. H2SO4). |
Electrode | Overpotential (V) | Tafel slope (mV dec−1) | Charge transfer resistance (Ω) |
---|---|---|---|
1 | 0.39 | 91 | 139 |
2 | 0.40 | 100 | 93 |
3 | 0.67 | 129 | 831 |
4 | 0.52 | 101 | 834 |
5 | 0.52 | 107 | 755 |
6 | 0.67 | 81 | 1143 |
GCE | >0.80 | 110 | >50![]() |
Pt | 0.06 | 26 | 1 |
GCE + Ni(OTf)2 | 0.67 | 126 | >20![]() |
The surface morphology of the modified electrodes was investigated using SEM before and after 1000 reductive cycles. Prior to reductive cycling, needle-like crystalline outgrowths with the morphology of Ni-230 are observed on the surface of 2, Fig. 4a and b. After 1000 CV cycles, a degradation in crystallinity is observed and the surface is more amorphous, Fig. 4c, although remnants of single crystals within the Nafion binder are visible upon magnification, Fig. 4d. The degradation of crystallinity following reductive cycling correlates with a significant reduction in overpotential and decreased charge transfer resistance. These improvements are attributed to the exposure of more active sites as the result of disruption of the surface by H2 bubbling during reductive cycling. Extended reductive cycling ultimately leads to a gradual increase in overpotential due to depletion of catalyst material during periods of vigorous hydrogen evolution. The SEM images of 1 before and after reductive cycling (Fig. S22†) are similar to those of 2, consistent with their similar changes in overpotential and charge transfer resistance upon reductive cycling. In contrast to 1 and 2, 4 and 5 exhibit amorphous surfaces prior to reductive cycling due to the decreased tendency of Ni-4 and Ni-5 to crystallize, Fig. S24 and S25.† After 1000 reductive cycles, the surfaces are more fractured, which correlates with enhanced HER activity. Notably, the surfaces of 3 and 6 show well-defined microcrystals that are largely unperturbed by reductive cycling, Fig. S23 and S26.† This can be attributed to the lower basicity of Ni-3 and Ni-6 resulting in less HER activity during reductive cycling. Consequently, reductive cycling has a negligible impact on the performance of 3 and 6.
![]() | ||
Fig. 4 SEM images of 2 (a, b) as drop cast on GCE and (c, d) after 1000 reductive cycles. Scale bars in a and c correspond to 100 μm and in b and d to 5 μm. |
Elemental mapping of the electrode surfaces by EDS before and after 1000 reductive cycles was performed to evaluate the effect on surface composition. Fig. 5a and b show the elemental mapping of nickel, nitrogen, and sulfur on the surface of 2 before and after 1000 reductive cycles, respectively. Each of the elements remain evenly distributed throughout the sample before and after reductive cycles. The EDS mapping plot depicted in Fig. 5c shows clearly defined nickel Lα and Kα peaks along with nitrogen and sulfur Kα peaks. Peaks for fluorine and oxygen from the Nafion binder can also be observed. A slight decrease in the intensity of the nickel peak after reductive cycling is observed, which is associated with the loss of catalyst from the electrode surface. EDS mapping of 1, 4, and 5 (Fig. S27, S29, and S30†) show a similar even distribution of nickel and other elements with a slight decrease in the peak intensity of nickel after 1000 reductive cycles. Interestingly, the EDS mapping plots of 3 and 6 (Fig. S28 and S31†), which do not show improvement in overpotential after reductive cycling, show no decrease in the peak intensity of nickel after reductive cycling. This is consistent with the observation in the SEM scan, where no changes in the surface morphology of the catalyst were observed even after 1000 reductive cycles.
In addition to the EDS mapping, XPS data were collected on 2 as a representative sample to verify the stability of the catalyst film after reductive cycling. The survey scans of 2 before and after 1000 reductive cycles are shown in Fig. S32.† Both scans display a prominent fluorine peak associated with the Nafion binder, along with nitrogen, sulfur, and nickel peaks from the catalyst. The similarities in the survey spectra before and after reductive cycling are consistent with retention of the catalyst structure. Fig. 6a shows the fitted peaks of the cationic nickel region. The high-resolution spectra of the core Ni 2p peak before cycling, Fig. 6a, shows a doublet at 854.6 eV and 871.82 eV (with satellite peaks) corresponding to 2p3/2 and 2p1/2, respectively, consistent with Ni2+ oxidation state. A comparison of high-resolution scans of the nickel region before and after reductive cycling, Fig. 6b, shows overlapping Ni 2p3/2 and Ni 2p1/2 peaks and satellites indicating no change in oxidation state.
![]() | ||
Fig. 6 XPS data for 2 showing (a) the fitted Ni2+ data before reductive cycling and (b) the Ni region before and after 1000 reductive cycles. |
To further confirm the stability of the catalyst, the XRD patterns of 1–3 were evaluated before and after reductive cycling. As shown in Fig. S34,† the pattern for 2 shows no change in peak positions after cycling, indicating the catalyst structure remains intact. A decrease in peak intensities was observed, which indicates a loss of the catalyst from the surface during reductive cycling. Similarly, comparative XRD diffraction patterns of 1 (Fig. S33†) show decreases in peak intensity after reductive cycling but no changes in peak positions. Interestingly, the XRD pattern of 3 (Fig. S35†) shows only minimal changes in peak intensities after reductive cycling with no change in peak position, supporting the SEM observation that the catalyst surface is not affected by reductive cycling.
Complex | Reduction potential (V vs. Fc) | pKa | Homogeneous overpotential (V vs. Fc) | Heterogeneous overpotential (V vs. RHE) |
---|---|---|---|---|
Ni-1 | −1.73 | 7.38 | 0.76 | 0.39 |
Ni-2 | −1.56 | 7.06 | 0.79 | 0.40 |
Ni-3 | −1.38 | <5 | 0.79 | 0.67 |
Ni-4 | −1.60 | 7.61 | 0.80 | 0.52 |
Ni-5 | −1.42 | 6.88 | 0.82 | 0.52 |
Ni-6 | −1.24 | <5 | 0.82 | 0.67 |
The Ni-1–Ni-6 catalysts were immobilized on GCEs to promote stability allowing for systematic variation of the HER overpotential under heterogeneous conditions. As highlighted in Table 2, electrodes prepared with Ni-3 and Ni-6 have the highest overpotential despite their accessible reduction potentials as the catalyst lack significantly basic sites for protonation. The HER activity of modified GCEs 1, 2, 4, and 5 was optimized by reductive cycling, which increased the number of active sites and reduced charge transfer resistance. Electrodes 1 and 2 had the lowest heterogenous HER overpotentials as they contain accessible thiosemicarbazonato groups. The introduction of bulky phenyl groups in the catalysts structure of 4 and 5 restrict H+ accessibility to the hydrazino nitrogen of the thiosemicarbazonato group increasing overpotential relative to 1 and 2. In contrast to most systems where the homogeneous HER method is good for systematic evaluation of ligand effects on the mechanism or HER activity and heterogeneous HER method is good for practical application of catalyst, this study is a rare example of a molecular catalysts system where immobilization is required for both systematic HER activity study and practical application.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5dt00005j |
This journal is © The Royal Society of Chemistry 2025 |