Biswanath
Das‡
*a,
Esteban A.
Toledo-Carrillo‡
b,
Guoqi
Li
c,
Jonas
Ståhle
a,
Thomas
Thersleff
d,
Jianhong
Chen
d,
Lin
Li
c,
Fei
Ye
b,
Adam
Slabon
e,
Mats
Göthelid
f,
Tsu-Chien
Weng
c,
Jodie A.
Yuwono
g,
Priyank V.
Kumar
g,
Oscar
Verho
*h,
Markus D.
Kärkäs
i,
Joydeep
Dutta
b and
Björn
Åkermark
*a
aDepartment of Organic Chemistry, Arrhenius Laboratory Stockholm University, Svante Arrhenius väg 16C, 10691 Stockholm, Sweden. E-mail: das.biswanath85@gmail.com; biswanath.das@su.se; bjorn.akermark@su.se
bDepartment of Applied Physics, KTH Royal Institute of Technology, Hannes Alfvéns väg 12, 114 19 Stockholm, Sweden
cSchool of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China
dDepartment of Materials and Environmental Chemistry, Stockholm University, Svante Arrhenius väg 16C, 10691 Stockholm, Sweden
eInorganic Chemistry, University of Wuppertal Gaußstr. 20, 42119 Wuppertal, Germany
fMaterial and Nanophysics, KTH Royal Institute of Technology, Hannes Alfvéns väg 12, 114 19 Stockholm, Sweden
gSchool of Chemical Engineering, University of New South Wales, Sydney 2052, Australia
hDepartment of Medicinal Chemistry, Biomedicinskt Centrum BMC, Uppsala University, SE-75123 Uppsala, Sweden. E-mail: oscar.verho@ilk.uu.se
iDepartment of Chemistry, KTH Royal Institute of Technology, SE-10044 Stockholm, Sweden
First published on 6th April 2023
The instability of molecular electrodes under oxidative/reductive conditions and insufficient understanding of the metal oxide-based systems have slowed down the progress of H2-based fuels. Efficient regeneration of the electrode's performance after prolonged use is another bottleneck of this research. This work represents the first example of a bifunctional and electrochemically regenerable molecular electrode which can be used for the unperturbed production of H2 from water. Pyridyl linkers with flexible arms (–CH2–CH2–) on modified fluorine-doped carbon cloth (FCC) were used to anchor a highly active ruthenium electrocatalyst [RuII(mcbp)(H2O)2] (1) [mcbp2− = 2,6-bis(1-methyl-4-(carboxylate)benzimidazol-2-yl)pyridine]. The pyridine unit of the linker replaces one of the water molecules of 1, which resulted in RuPFCC (ruthenium electrocatalyst anchored on –CH2–CH2–pyridine modified FCC), a high-performing electrode for oxygen evolution reaction [OER, overpotential of ∼215 mV] as well as hydrogen evolution reaction (HER, overpotential of ∼330 mV) at pH 7. A current density of ∼8 mA cm−2 at 2.06 V (vs. RHE) and ∼−6 mA cm−2 at −0.84 V (vs. RHE) with only 0.04 wt% loading of ruthenium was obtained. OER turnover of >7.4 × 103 at 1.81 V in 48 h and HER turnover of >3.6 × 103 at −0.79 V in 3 h were calculated. The activity of the OER anode after 48 h use could be electrochemically regenerated to ∼98% of its original activity while it serves as a HE cathode (evolving hydrogen) for 8 h. This electrode design can also be used for developing ultra-stable molecular electrodes with exciting electrochemical regeneration features, for other proton-dependent electrochemical processes.
Molecular catalysts offer an attractive alternative to metal oxide and hydroxide-based electrodes due to their improved product selectivity, and the possibility of systematic improvement.9,10,14 As has been demonstrated lately, the instability of the molecular electrocatalysts, under oxidative–reductive environments, can be overcome by immobilizing or anchoring them onto conducting solid surfaces through covalent bonding or π–π stacking interactions.15–19 Ruthenium-based complexes have so far shown the best performance in terms of turnover number (TON), turnover frequency (TOF), and overpotential.17,20–22 However, for commercial applications, further improvement of their stability, as well as strategies for catalyst regeneration are needed.
Fluorine-doped carbon cloth (FCC) has been reported and recently tested by us a as robust and durable conductive support that allows for straightforward structural modification and catalyst grafting.23,24 We recently reported the preparation of CoPFCC electrode consisting of {CoII(mcbp)(H2O)} anchored onto pyridine-modified FCC, which showed high stability and robust features under WO conditions.24 However, to enable sustainable production of H2 from water, further improvement of overpotential and TO (turnover) values is needed. We were only able to regenerate nearly about 78% of the original WO activity with the CoPFCC electrode following its first use (for 72 h) as an anode, which highlights further scopes for upgrading. The CoPFCC electrode could not show any appreciable proton reduction features, representing the need for another electrode (cathode) for proton reduction to avail H2 from water electrolysis.
In recent years, high performing dual functional electrodes that can be used for both OER and HER, have attracted attention due the possibility of using them as both anode and cathode in efficient electrolysis cells for overall water electrolysis.25–30 Usually the HER activity increases with decreasing the pH, while OER is accelerated at higher pH. Therefore, developing electrodes/electromaterials that simultaneously can perform both HER and OER at neutral pH is an additional challenge.31 Aiming towards a bifunctional and fully regenerable electrode setup for unhindered production of H2 from water, we decided to investigate the anchoring of an active ruthenium containing catalyst onto the pyridine (–CH2–CH2–pyridine) modified FCC (PFCC) surface. Ruthenium-containing catalysts have been extensively reported to show exciting features in both oxidations and reductions.32–37 Moreover, ruthenium(II) is known to form strong covalent bonds with N-donor ligands.9,17,22,38,39 The favourable Ru(II)–N (of pyridine from PFCC) interaction was envisioned to replace one of the aqua molecules in catalyst 1, an analogue of RuII(mcbp)(OH2)(py)2 which has previously been reported as a high-performing catalyst for electrochemical WO (TOFmax of 40000 s−1 for WO at pH 9).22
The mcbp2− framework, with its two benzimidazole carboxylate units, acts as a non-innocent ligand that can facilitate the proton-coupled electron transfer (PCET) processes. RuPFCC (ruthenium catalyst anchored onto PFCC) was designed to combine the electrocatalytic activity from the molecular counterpart (of 1), with the current density, and robust features of the conductive support.9,17,20–22,35,40 The pyridines were covalently attached to the carbon cloth separated by two –CH2 units, to allow for optimal distances for electron transfer and charge recombination.41
Importantly, we found that the overpotential for OER is significantly reduced at Ru-2H2O (an overpotential of 0.33 eV) compared to that at Ru-1H2O (0.67 eV). Further, we found that the rate-limiting step is *OOH to O2 evolution for Ru-1H2O, while it is *OH to *O formation for Ru-2H2O. Our HER calculations showed that the free energy of H adsorption (H*) is pushed closer to thermoneutral for Ru-2H2O compared to that at Ru-1H2O , suggesting that it is also easier for HER to proceed at Ru-2H2O.
Finding this additional HER capability, i.e., understanding the dual functionality of [RuII(mcbp)(H2O)2] is one of the novelties of this work. In terms of Ru–N coordination and bond length variations, 1 followed very similar reactive intermediate structures as discussed by us with [RuII(mcbp)(py)2] catalyzed oxygen evolution.22 It is important to note that an increase in the number of water molecules attached to the active site makes the ligand (mcbp2−) less firmly bound and the catalyst becomes more susceptible to decomposition while turning over. Anchoring the catalyst onto a conductive solid surface through a suitable linker can improve both the stability and the overall reactivity, leading to the enhancement of the current density.17,20
FCC and PFCC showed very similar high-resolution XPS spectra (C 1s and Ru 3d, Fig. 2d, S10†) with two peaks at about 284 and 292 eV, which arises from sp2 carbons in the graphitic structure and the CF2 groups incorporated upon fluorination, respectively.42 In contrast, RuPFCC shows a higher peak intensity of the band located at 284 eV and an extra Ru 3d signal, which is a composition of three different Ru peaks, attributed to N(pyridine)-Ru, N(mcbp ligand)-Ru, and O-Ru.42 Impedance results are shown as Nyquist plots (Fig. 3f), where the semicircles in the high-frequency regime are assigned to a charge transfer process (Mechanism I), a linear region in the intermediate frequencies is associated to mass transport processes (Mechanism II), and a final linear region with a higher slope in the low frequency region is assigned to capacitive behaviour (Mechanism III). The diameter of the semicircle corresponds to the charge transfer resistance (Rct).43,44 A clear reduction in the Rct is observed for RuPFCC, demonstrating an enhancement in the electrochemical activity. The appearance of capacitive behavior at higher frequencies for RuPFCC, is associated to a faster mass transport to the metal centers. In other words, the active sites in the molecular catalyst are easily accessible for substrate (H2O) molecules.
Fig. 3 SEM images for (a) FCC (b) PFCC and (c) RuPFCC. (d) XPS spectra of C 1s of FCC, PFCC and RuPFCC, (e) TGA of CC, FCC, PFCC and RuPFCC (discussion in ESI†), (f) EIS of PFCC and RuPFCC, (g) CVs (−1.05 V to +2.2 V, vs. RHE) of a pH 7 phosphate buffer (0.1 M) solution using 1 cm × 0.5 cm RuPFCC (green trace), FCC (red trace), PFCC (blue trace) and a glassy carbon (d = 1 mm) (grey trace) as the working electrodes (WE) at scan rate of 100 mV s−1. Pt wire and Ag/AgCl (3.0 M KCl) electrodes were used as counter electrode and reference electrode respectively, (h) CVs (0.35 V to +2.2 V, vs. RHE) using 1 cm × 0.5 cm RuPFCC (green trace), FCC (red trace), and a glassy carbon (d = 1 mm) (grey trace) as the working electrodes (WE), (i) CPE results focusing over first 3.4k seconds performed at 1.71 V [using 1 cm × 0.5 cm RuPFCC (green)] and at 1.81 V and 1.71 V [using 1 cm × 0.5 cm FCC (blue and red trace respectively)]. |
Controlled potential electrolysis (CPE) experiments were performed at 1.71 V (8 h and 20 h), and −0.81 V to confirm WO (through oxygen detection) and WR (through hydrogen detection) and to investigate electrode performance (charge accumulation and current density) and stability. No drastic changes in the charge or current were observed, however, a continuous electron transfer could be seen in both the cases. For WO, RuPFCC yielded a TO of >1.7 × 103 over a period of 8 h at 1.71 V. During the first two hours of the first CPE experiment, 12.5 μmol of oxygen was evolved with 82% of Faraday efficiency (Table 1, entry 1 and Fig. S15†). Yagi et al., recently reported 91% Faraday efficiency of oxygen evolution with the RuCP electrode {RuCP = immobilized proximal,proximal-[Ru2L(C8Otpy)2(OH)(OH2)]3+ on carbon paper, where C8Otpy = 4′-octyloxy-2,2′; 6′,2′′-terpyridine and L = 5-phenyl-2,8-di(2-pyridyl)-1,9,10-anthyridine} for the first one hour of electrolysis at 2.01 V at pH 7.20 The Faraday efficiency with RuCP electrode over the first 3 h decreases to 45% when the applied potential was decreased by 450 mV. In our study, when the CPE using RuPFCC was repeated at 1.81 V for 48 h a TON of >7.4 × 103 could be achieved. The efficiency of the resulting electrode after CPE experiments were checked before and after washing with water in a fresh buffer (pH 7) solution. When the electrode was reused as a WO anode under the same reaction conditions after 48 h of continuous oxygen evolution, it evolved 3.5 μmol of oxygen over the first two hours of CPE with a 44% Faraday efficiency (Table 1, entry 2). Compared to the first CPE, a clear difference (Fig. 4a) in the current density was also observed this time.
Fig. 4 (a) Positive charge (Q) vs. time (s) plot for CPE over 8 h, at 1.71 V (green trace, using 1 cm × 0.5 cm RuPFCC as working electrode), next day over 2 h (red trace) and after the electrochemical regeneration of the electrode (ERE, see ESI†) over 1 h. Inset shows corresponding current (mA) vs. time (s) plot. For first 1k seconds 98% recovery was possible, (b) overall oxidation and ERE (electrochemical regeneration experiment) timeline; (c) CVs (from +0.35 V to +2.21 V) before starting CPE experiment (dark green), after 20 h (black trace) and 48 h (purple trace) of CPE experiments at +1.81 V. Red and blue traces show regeneration of the electrocatalytic activity after 68 h of oxidation; (d) CVs (from –1.0 V to +2.21 V) of RuPFCC before CPE experiment (dark green), after 16 days (red) kept in the same buffer solution under N2 and multiple CPE experiments (4 oxidations of 4 h, 8 h, 20 h and 48 h at +1.81 V and 3 ERE of 30 min, 1 h and 2.5 h at –0.59 V) and after last ERE (6 h) (blue trace); (e) potential (V) vs. time (min) (above) and corresponding H2 detection (below) plots of a chronopotentiometric experiment at a constant current of –5 mA cm−2 in pH 7 phosphate buffer electrolyte over 2 h and (f) electrochemical response of RuPFCC under the HER conditions. In both the HER related experiments, RuPFCC (1 cm × 0.5 cm) after 8 h of WO was used as working electrode. Carbon paper (1 cm × 0.5 cm) and Ag/AgCl (3.0 M KCl) electrodes were used as counter electrode and reference electrode respectively. |
A small positive shift in WO oxidation potential and a decrease in the current density were observed after every oxidation (CPE) process (Fig. 4c and d). A longer CPE process caused a more prominent change. 1H and 19F NMR (500 MHz) of the solution after the electrolysis confirmed that no catalyst and fluoride leaching happened during the CPE process. In an additional experiment, the Pt wire was replaced by a carbon rod as counter electrode for 48 h to investigate whether Pt corrosion is impacting the overall reactivity. No appreciable difference (<5% at 1.75 V) in the cyclic voltammetric response was observed.
In a separate experiment, dissolved oxygen was monitored by chronopotentiometry keeping a constant current density at 2.6 mA cm−2 (Fig. S16†). A sharp increase of oxygen close to the working electrode at a potential of 1.44 V (oxygen evolution rate of 1.048 μmol min−1 L−1 for first 30 min) was seen, corresponding to an overpotential of ∼215 mV. This indicates an appreciable shift (∼415 mV) in the overpotential of WO upon coordination of the electrocatalyst onto the PFCC through the pyridine linkers.39
Entry | Catalyst | Conductive support | pH | η (mV) | TO (h, AP) |
---|---|---|---|---|---|
a Calculated value from the reported electrochemical response. | |||||
1 | [Ru(mcbp)(H2O)2] (1) | PFCC (this work) | 7 | 215 | 7430 (48, 1.40) |
2 (ref. 20) | Ru2L(C8Otpy)2(OH)(OH2)]+ | Carbon paper | 7 | 260 | 4080 (35, 1.64) |
3 (ref. 46) | Ru2L(cptpy)2(OH)(OH2)]+ | Nano-TiO2/ITO | 7 | 530 | 416 (3, 1.60) |
4 (ref. 46) | [Ru2(Hcptpy)2L(μ-Cl)]3+ | Nano-TiO2/ITO | 7 | 830a | 22 (3, 1.60) |
5 (ref. 17) | Ru(tda)(4, 4′-bpy)]15(4, 4′-bpy) | MWCNT/GC | 7 | 400 | 2 × 105 (12, 1.86) |
6 (ref. 39) | Poly-[Ru(tda)(pyrS)2] | Carbon paper | 7 | 385a | 5000 (0.5, 1.40) |
An increase in the broad irreversible reduction wave was observed at around −0.19 V after 72 h of continuous OER at 1.71 V, which can be attributed to the pyridinium(py+)/pyridine(py) redox wave, (Fig. S17†).31 This disappeared after 2 h of ERE at −0.59 V (Fig. S17†). Moreover, pH changes of the bulk solution from 7.1 to 6.6 after 72 h of continuous OE was observed. An increase in the protonation of the free pyridyl units of the RuPFCC surface upon prolonged WO experiments leads to a decrease of the electrocatalytic WO activity due to the lesser availability of the proton acceptors on the electrode surface. These finding allowed us to propose a mechanism (Figure S18†) where the ERE experiment allows deprotonation of the pyridinium units and regeneration of the WO efficiency.
The electrode potential reached a stable value of −0.79 V vs. RHE after 55 min, showing no sign of degradation (Fig. 4e). The hydrogen content reached a stable value after 70 min, which could be associated to the release of hydrogen into the gas phase. Another set CPE experiments at −0.79 V (vs. RHE) reveals that RuPFCC loses its electrocatalytic proton reduction activity by 8% (measured at −1.0 V) after 3 h of constant PR and by 24% after 20 h of PR (Fig. 4f). During the 3 h PR experiment at applied potential of −0.79 V, a total TOPR of 3680 was calculated from the current density plot (see ESI†). In another experiment, when a freshly prepared RuPFCC was tested under HER conditions, at −0.79 V, it showed nearly 18% lower current density in the first 2 h of the experiment (Fig. S19†) than RuPFCCox. This supports our proposed mechanism (Fig. S18†) where the formation of the pyridinium units on RuPFCC surface (RuPFCCox) during the WO experiment makes it a better electrode for HER. While releasing H2 during HER process, pyridinium units transform back to pyridines which makes it again suitable for WO, i.e., regeneration. This simple but effective design strategy can potentially be used for electrode preparation in various proton dependent electrochemical processes.
The overall water electrolysis at pH 7 and quantification of the evolved gas from the anode and cathode compartments were performed using a commercial H-cell and a home-made volumetric extension (Fig. S15†). It confirmed the anodic production of O2 and cathodic production of H2 while using RuPFCC as electrodes in both the compartments. An overall cell voltage of 1.78 V at pH 7 for anodic OER (starts at 1.44 V) and cathodic HER (starts at −0.34 V) was calculated. Although these values (overpotential and overall cell voltage) are on the higher side compared to reported state-of-the-art RuO2/IrO2 cluster and nanostructured transition metal electrocatalysts (overall cell voltages are around 1.55 V), low loading (∼0.04 wt% ruthenium) of the ruthenium centres on RuPFCC, together with its regeneration feature can make the water electrolysis cost-effective, as well environmentally more benign.49,50
Electron microscopy studies were carried out to examine the chemical stability of the RuPFCC electrode after continuous electrolysis for 12 h of OER (∼98% regenerable), 16 days (∼74% regenerable), 3 and 4 months (<60% regenerable), of multiple OER and HER experiments. It can be observed that the extent of over-oxidation of the molecular catalysts gradually increases with the number of OER and HER. As indicated in the TEM images (Fig. S21–S23†), the oxidation product, RuO2 nanoparticles were formed on the molecular electrode and the amount increases with the time of electrolysis.
As previously reported by our group, to be effective, the ruthenium center of 1 must proceed RuIV/RuV oxidation, which is involved in the rate-limiting oxygen–oxygen bond formation step through water nucleophilic attack.22 For this, it is pivotal to have the mcbp2− type ligand system remain attached to the ruthenium center. Over-oxidation can produce different kinds of ruthenium oxides, among which (i) mesoporous RuO2 (MP-RuO2), (ii) commercially available RuO2 nanoparticles (C-RuO2) and (iii) high surface area containing RuO2, also known as Ad-RuO2 are extensively discussed.51,52 While MP-RuO2 showed interesting OER activities, others have shown less to almost no reactivity. Most likely, the molecular catalysts on PFCC after more than 65 h of continuous OER experiment transform into appreciable amount of nonreactive/unreactive ruthenium oxides that reflects in incomplete (<95%) regeneration of the electrode's activity. The formed RuO2 nanoparticles on the PFCC surface seem to be less active than 1. Although no catalyst leaching could be observed (by high resolution NMR with 10% D2O insert) in the resulting electrolyte solution after more than 65 h of continuous electrolysis, after 4 months of multiple OER and HER experiments fragmented H2mcbp ligands (Fig. S25†) could be seen, indicating a slow decomposition of the electrocatalyst.
The morphology and roughness of PFCC and RuPFCC were determined using SEM using Zeiss GEMINI® Ultra 55. Transmission electron microscopy analysis was performed using an aberration-corrected Themis Z instrument from Thermo Fisher. TEM was operated at 300 kV in Scanning TEM (STEM) mode. A TEM grid was prepared by immersing RuPFCC in 1 M aqueous KOH and ultrasonicating it for 15 minutes, after which it was neutralized with 1 M HCl and dispersed onto a thin carbon support film. This allowed detaching some of the aggregates from RuPFCC (Fig. 3c), enabling a detailed TEM analysis (Fig. S9†). XPS was performed using Thermo Scientific ESCALAB 250Xi with a monochromatic Al Kα source (hν = 1486.6 eV). The Thermo Scientific™ Avantage software was used for data analysis and curve fitting. Raman spectral analysis was performed with an iHR550 Spectrometer (Horiba Jobin Yvon) in the range between 200 to 4000 cm−1 with a 633 nm laser light for excitation. A grating of 1200, a pinhole of 130 with 200 acquisition and a time of 1 s were used to see the changes between FCC, PFCC, and RuPFCC (Fig. S10†). TGA was performed using a TGA Q500 equipment (TA Instrument) under airflow to determine weight losses due to the elimination of surface functionalities and degradation of carbon material. Electrochemical measurements were performed in a conventional three-electrode setup using a Potentiostat/Galvanostat/ZRA Gamry Interface 1010E. EIS was performed in potentiostatic mode at open circuit potential (bias potential) within a frequency range from 100 kHz to 10 MHz. The amplitude of the sinusoidal was 10 mV rms.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ta00071k |
‡ BD and ET equally contributed to the experimental work. |
This journal is © The Royal Society of Chemistry 2023 |