Open Access Article
Gabriella P. Bein
,
Sergio Fernández
,
Stephen J. Tereniak
,
Renato N. Sampaio,
Alexander J. M. Miller
* and
Jillian L. Dempsey
*
Department of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290, USA. E-mail: dempseyj@email.unc.edu; ajmm@email.unc.edu
First published on 5th November 2025
A first-row transition metal catalyst, [Fe(tpy)(Mebim-py)(NCCH3)]2+ (tpy = 2,2′:6′,2′′-terpyridine, Mebim-py = 1-methylbenzimidazol-2-ylidene-3-(2′-pyridine)) mediates CO2 reduction to CO at passivated p-Si photoelectrodes with applied potentials 240 mV positive of the standard CO2/CO reduction potential. The molecular catalyst's selectivity for CO was retained under photoelectrochemical conditions, with negligible direct proton reduction promoted by the photoelectrode. The faradaic efficiency for CO (44 ± 6%) was slightly enhanced relative to the catalyst performance in the dark (33%). A photosynthetic cell based on this photocathode system, coupled with ferrocene oxidation at the anode, successfully operated at a cell voltage of −1.2 V. The photovoltage generated by illumination of p-Si–CH3 met and surpassed the potential required for CO2 reduction when coupled with ferrocene oxidation. By leveraging a low-overpotential CO2 reduction electrocatalyst, a photo-assisted electrochemical efficiency of 0.15% and applied bias photon-to-current efficiency of 0.05% was achieved for this single-junction cell, ultimately storing 46 kJ mol−1 (11 kcal mol−1) of photon energy.
One class of photoelectrochemical cells for artificial photosynthesis integrates molecular catalysts dissolved in the electrolyte solution that are selective for carbon dioxide reduction with visible light absorbing semiconductors.7–9 Limitations of this architecture include competition between the CO2 reduction reaction at the molecular catalyst and photocathode surface-based redox processes, such as photocathode corrosion and hydrogen evolution.10–12 However, passivation of p-type silicon photoelectrodes with polymeric or oxide coatings,13,14 or by direct covalent modifications of the Si lattice, has been shown to be effective at improving the photoelectrode stability to surface oxidation and minimizing unwanted side reactivity.15–18 These advances, together with the advantages of solution-based catalysts towards well-behaved electrochemical responses, motivate further work in this architecture. Recently, we reported that methyl-terminated p-Si photoelectrodes (p-Si–CH3) interfaced with a homogeneous ruthenium polypyridyl catalyst, [Ru(tpy)(Mebim-py)(NCCH3)]2+ (tpy = 2,2′:6′,2′′-terpyridine, Mebim-py = 1-methylbenzimidazol-2-ylidene-3-(2′-pyridine)), in the electrolyte solution can selectively drive CO2 reduction without parasitic hydrogen evolution from the photoelectrode. Unlike unpassivated p-Si–H, which rapidly becomes charge transfer resistant as the surface oxidizes, the p-Si–CH3 was stable under photoelectrolysis conditions and sustained a steady photocurrent. The Ru catalyst was able to access the same intrinsic product selectivity for CO at the p-Si–CH3 photoelectrode as was observed at metallic electrodes, but with a significant photovoltage of 460 mV.19 Importantly, the selectivity for the CO2 reduction reaction was rationalized to occur because the reduction of CO2 mediated by the ruthenium catalyst kinetically outcompetes direct proton reduction at the photocathode, revealing an important design principle for selective fuel formation.
One disadvantage of the [Ru(tpy)(Mebim-py)(NCCH3)]2+ system is that this complex has a large overpotential, η, (eqn (1)).
![]() | (1) |
The catalyst requires an applied potential significantly beyond the standard reduction potential for the CO2 reduction reaction to form the active catalyst species, with the reduction of the ruthenium species to the active [Ru(tpy)(Mebim-py)]0 species occurring at E1/2 = −1.94 V vs. Fc+/0.20 Under the conditions at which the catalyst operates (95
:
5 CH3CN
:
H2O solvent mixture), the standard reduction potential for CO2 reduction to CO is −1.44 V vs. Fc+/0.21 Even when paired with an anode reaction that operates with no overpotential, the photovoltage generated at the p-Si–CH3 photoelectrode was insufficient to offset an overpotential of this magnitude and a relatively negative applied potential was still needed for photoelectrochemical CO2 reduction, such that the system is photocatalytic (ΔG < 0), not photosynthetic (ΔG > 0).6 Furthermore, the full cell potential and efficiency, including both the cathodic and anodic half-reactions, as well as full system losses, must be considered in order to achieve photon energy storage in solar fuels.
We hypothesized that an energy-storing photosynthetic system could be achieved if the same robust silicon photocathode operating with a large photovoltage was paired with a molecular catalyst that operates in the dark with a minimal overpotential. In this case, CO2 reduction at the cathode could proceed at an effective “underpotential”, defined as a negative overpotential. In this situation, the applied electrochemical potential is less than the standard reaction potential for the reaction of CO2 to CO, and photons supply the remaining thermodynamically required energy. Beyond the impact of demonstrating photon energy storage, an example of energy-storing photosynthesis would provide an opportunity for standardizing efficiency assessment protocols, motiving our work.
To examine the hypothesis that a low-overpotential catalyst could enable energy-storing photosynthesis in conjunction with a single-junction p-Si photocathode, we identified the [Fe(tpy)(Mebim-py)(NCCH3)]2+ CO2 reduction catalyst (Fig. 1), which mediates CO2 reduction to CO by a similar mechanism to the analogous Ru complex, but operates at a substantially reduced overpotential for CO2 reduction (η = 150 mV).22 Combining this low-overpotential catalyst based on an abundant first-row transition metal element with p-Si–CH3, a durable and well-passivated visible-light-absorbing photocathode based on the industry standard photovoltaic material, we access a photocathode system competent for CO2 reduction with faradaic efficiencies (FEs) exceeding those achieved with a metallic electrode. Pairing this photocathode with ferrocene oxidation at a dark anode affords a photosynthetic system competent for CO2 reduction. While many works focus solely on the CO2 reduction reaction that produces the fuel of interest, it is important to carefully consider both half-reactions when designing a photosynthetic cell in order to establish energy-storing properties. Operating at an applied cell potential of −1.2 V under 1 sun illumination, 46 kJ mol−1 (11 kcal mol−1) of photon energy are stored in the CO fuel product. This demonstration is one of the first known examples of true energy-storing photosynthesis based on a molecular catalyst operating with a single-junction semiconductor and provides a blueprint for advancing photon-to-fuel efficiency in molecular catalyst-based photoelectrosynthesis systems.
Under a CO2 atmosphere, we observe loss of chemical reversibility of the Fe2+/0 wave and an enhancement of the cathodic current at the p-Si–CH3 (Fig. 3). This peak-shaped wave is interpreted as a four-electron, two-proton reduction of the FeII species coupled to CO2 reduction to form the CO-bound Fe0 species, Fe(tpy)(Mebim-py)(CO), based on reactivity observed in the dark at metallic electrodes.22 CO release enables catalyst turnover, but the slow kinetics for this catalyst (TOF = 8 s−1) manifest in minimal current enhancement on the cyclic voltammetry timescale. A new oxidation feature at −0.80 V vs. Fc+/0 is apparent in the return sweep, heralding the formation of Fe(tpy)(Mebim-py)(CO), which has a characteristic oxidation at this potential.22 These voltammograms indicate that the Fe catalyst is sufficiently photostable under these conditions and can mediate CO2 reduction at very mild applied potentials with a p-Si–CH3 photoelectrode under illumination, notable since the photocatalytic activity is observed at potentials more positive than the standard electrode potential for CO2 to CO in the dark (
vs. Fc+/0 in 95
:
5 CH3CN
:
H2O solvent mixture).21 There is a trade-off between the catalyst overpotential and its activity, as seen in how the peak catalytic photocurrent (ip,c) for [Fe(tpy)(Mebim-py)(NCCH3)]2+ at 0.1 V s−1 is substantially lower than that recorded for [Ru(tpy)(Mebim-py)(NCCH3)]2+ under the same conditions (Fig. S2). This observation is consistent with the relative current response of these two catalysts observed at glassy carbon electrodes,22,23 which is in line with the three order-of-magnitude difference in their respective turnover frequencies. It is significant that the photoelectrochemical response at p-Si–CH3 is well-behaved and closely resembles the voltametric waveforms at glassy carbon; the well-defined and interpretable cyclic voltammograms contrast the commonly observed broad and low current amplitude redox events of p-Si photoelectrodes modified with molecular catalysts on the surface.24–27 Furthermore, the comparable photovoltages achieved for the Fe and Ru systems exceeded expectations, as p-Si photovoltage is known to correlate with E°′ of the redox-active analyte in solution until inversion conditions are reached.28
After a 3600 s CPPE, during which −0.374 C was passed, the headspace of the reaction vessel was analysed by gas chromatography for gaseous product detection. CO was detected with a FE of 52% and no H2 was detected. The FECO value is slightly higher than that reported for this catalyst at glassy carbon electrodes under static headspace conditions in the dark (33%). Based on mechanistic investigations in previous studies of this catalyst, we attribute the remainder of the charge passed to the build-up of the Fe(tpy)(Mebim-py)(CO) intermediate in solution.22 Voltammograms recorded immediately before and after CPPE are shown in Fig. S4. The p-Si–CH3 photoelectrode was durable under the photoelectrocatalytic conditions, evidenced by the lack of competitive HER activity and by the sustained ability to pass charge, as observed previously.19 With AM1.5G illumination at −1.31 V vs. Fc+/0, the average across three trials under the same conditions were FECO = 44 ± 6% and FEH2 = 1 ± 1%. The product selectivity was also reproducible at various applied potentials and with different illumination sources, with FECO ranging from 33–76%, depending on the conditions (Table S1).
Traditionally, CO2 reduction electrolysers have coupled the reductive half-reaction with oxygen evolution from water (OER) at the anode.30 However, depending on the catalyst employed, water oxidation often operates with a high overpotential because of slow reaction kinetics,31,32 consuming up to 90% of the total cell voltage.33 In some reports, OER is replaced with organic alcohol oxidation at the anode to in order to operate the photosynthesis cell at moderate cell voltages and improve energy conversion efficiency.33–37 Here, we selected ferrocene oxidation as a convenient and well-behaved anode half-reaction with a well-defined potential (E1/2 (Fc+/0) = 0 V vs. Fc+/0) that can operate at essentially zero overpotential, leading to the overall reaction of eqn (2) for our proof-of-concept demonstration. A two-electrode CPPE was performed with 1 mM [Fe(tpy)(Mebim-py)(NCCH3)]2+ in the working compartment and 10 mM Fc and 10 mM [Fc][PF6] in the counter electrode compartment. The 1
:
1 Fc
:
Fc+ was chosen for the counter compartment to approach equilibrium conditions for eqn (2) and to aid in measuring the change in open circuit potential (OCP) of the counter compartment during the photoelectrolysis. The cell resistance measured between the cathode and anode across the dividing glass frit of the H-cell was ca. 1 kΩ, showing significant potential drop as a result of the photosynthetic cell geometry.
![]() | (2) |
Under AM1.5G 100 mW cm−2 simulated solar illumination, a cell potential (Ecell) of −1.2 V was applied for 1 h between the p-Si–CH3 photocathode and the Pt mesh anode (Fig. 5). Under illumination, this cell voltage was sufficient to drive the reaction forward, passing on average 0.276 C of charge and producing CO with 56 ± 18% FE (Table S2). The OCP of the counter compartment shifted to more positive potentials, consistent with the oxidation of ferrocene to ferrocenium. The production of Fc+ at Pt mesh had near unity FE (96 ± 2%, see Table S3).
The Gibbs free energy (ΔG) of the net cell reaction defined in eqn (2) can be used to quantify how much light energy is stored in the fuel:
![]() | (3) |
485 C mol−1). The applied voltage Ecell is 240 mV less negative than the potential associated with the CO-producing reaction of eqn (2) at the cathode (
vs. Fc+/0 in 95
:
5 CH3CN
:
H2O solvent mixture) with ferrocene oxidation at the anode.21
With Ecell = −1.2 V, we find ΔG = 46 kJ mol−1 (11 kcal mol−1). As the reaction is endergonic under state conditions, the photosynthetic cell requires the incident light energy to drive this reaction, ultimately storing 46 kJ mol−1 of photon energy in the CO fuel product.6
The photo-assisted electrochemical efficiency (ηPAE) of this photosynthetic cell is the ratio of the system output to total input power:38,39
![]() | (4) |
An overall ηPAE of 0.15% is quantified for this photosynthetic cell operating at −1.2 V. In context, the ηPAE quantified here is on the low end compared to heterogeneous photoelectrochemical CO2 reduction systems. While ηPAE is rarely quantified in photoelectrochemical CO2 reduction studies, there are reports for heterogeneous systems with ηPAE ranging from 0.37% to 11%.40–43 To the best of our knowledge, there are no other reports to date of ηPAE for a molecular catalyst driving photoelectrochemical CO2 reduction catalysis.
With this data we can also calculate the related applied bias photon-to-current efficiency (ABPE) metric. By nature of utilizing a single-junction p-Si photoelectrode, an applied bias is required to supplement the photovoltage and drive CO production. ABPE represents the net chemical power output relative to the incident solar power:38,39,44
![]() | (5) |
Applying eqn (5), the ABPE is 0.05%. This metric is commensurate with work by Li and co-workers, who reported a similar ABPE of 0.06% for a cobalt polypyridyl catalyst attached to carbon nanotubes on a TiO2-coated Si photoelectrode.45 While some photo-assisted electrochemical efficiency loss comes from the FE of the cathodic reaction (which can be improved by operating under constant CO2 flow),22 the major contributing factor to both low ηPAE and low ABPE is likely the low photocurrent, which we hypothesize is limited by low catalyst turnover frequency and by low internal quantum efficiency (IQE) of the photocathode. Slow catalyst kinetics lead to efficiency-limiting carrier recombination when the catalyst is not available to accept photogenerated charge carriers. A faster catalyst could increase the photon-to-current efficiency by keeping pace with the incident photon flux. The IQE is affected by charge separation, charge carrier recombination, and the heterogeneous electron transfer rate.19,46 Indeed, previous studies of the [Ru(tpy)(Mebim-py)(NCCH3)]2+-mediated CO2 reduction at p-Si–CH3 showed that photocurrent was limited by the IQE.19 IQE can be improved by increasing electron-transfer rate constants and with the use of high-quality Si wafers. These mechanisms for increasing jph provide a roadmap towards improving the efficiency metrics, ηPAE and ABPE. Further, despite the low ηPAE and ABPE values, this work demonstrates exciting proof-of-concept for storing incident photon energy as fuel in photosynthetic system with a molecular catalyst-based photocathode operating with a single-junction semiconductor light harvester.
| This journal is © The Royal Society of Chemistry 2025 |