Next Article in Journal
Nitrogen-Rich Tetrazole–Amide-Functionalised Zn Metal–Organic Framework as Catalyst for Efficient Catalytic CO2 Cycloaddition with Epoxides
Previous Article in Journal
Conversion of CO2 into Glycolic Acid: A Review of Main Steps and Future Challenges
Previous Article in Special Issue
Development and Optimization of Air-Electrodes for Rechargeable Zn–Air Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrocatalytic Hydrogen Evolution of Transition Metal (Fe, Co and Cu)–Corrole Complexes Bearing an Imidazole Group

1
Guangdong Provincial Key Laboratory of Fuel Cell Technology, School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510641, China
2
School of Materials Science and Energy, Foshan University, Foshan 528000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(1), 5; https://doi.org/10.3390/catal14010005 (registering DOI)
Submission received: 4 November 2023 / Revised: 7 December 2023 / Accepted: 14 December 2023 / Published: 19 December 2023

Abstract

:
The study of the hydrogen evolution reaction (HER) by non-noble transition metals is of great significance for the production of hydrogen energy. In this work, a new 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole) phenyl]-corrole and its metal complexes (metal = Co, Cu, Fe) were synthesized and used for electrocatalyzed HER in DMF organic solvent and aqueous media. The prepared cobalt corrole showed the best catalytic performance in both media. Its turnover frequency (TOF) and catalytic efficiency (C.E) could reach 265 s−1 and 1.04 when TsOH was used as the proton source in a DMF solvent. In aqueous media, its TOF could also reach 405 h−1. The catalytic HER may go through an EECEC or ECEC (E: electron transfer, C: chemical step) pathway for these catalysts, depending on the acidity and concentration of the proton source. The present work successfully demonstrates that imidazole at a meso-phenyl group may improve the electrocatalytic HER activity of transition metal corroles.

Graphical Abstract

1. Introduction

With the continuous consumption of non-renewable fossil fuels and environmental pollution caused by their combustion, exploring new sustainable and environmentally friendly energy has become the focus of scientific research. Hydrogen energy is the most promising renewable energy because of its high calorific value and pollution-free combustion [1,2]. A variety of methods have been reported for producing hydrogen, with the electrocatalytic hydrogen evolution reaction (HER) being particularly significant [3,4,5]. Due to the slow reaction kinetic of HER, the reaction usually needs to be facilitated by the addition of a catalyst. Platinum-based catalysts have excellent catalytic performance for HER; however, their high cost and scarcity limit their wider application [6,7,8]. Therefore, exploring earth-abundant catalysts that effectively catalyze HER is crucial.
Currently, many low-cost transition metals have been used in catalytic HER studies, including cobalt [9,10], copper [11,12,13], nickel [14], iron [15], etc. Macrocyclic molecular catalysts have also been used for HER due to their unique chemical structure and tunable activity [16,17,18]. Metal porphyrins have been extensively used in elelctrocatalytic HER [19,20,21]. Corrole is a class of trianionic macrocyclic ligand in the porphyrin family. It has the capacity to stabilize encapsulated metal with higher oxidation [22]. This is conducive to electrocatalytic hydrogen evolution by lowering the overpotential [23]. Recently, lots of studies have focused on metal–corrole electrocatalysts in HER [21,24,25]. The catalytic properties of metal corroles may be modulated by modifying the groups at their meso-positions [26]. It has been found that the strong electron-withdrawing groups such as pentafluorophenyl [27], cyanophenyl [28], and nitrobenzene [29] could enhance the catalytic HER ability of metal–corrole catalysts. In addition, the introduction of proton relay groups such as amine [30], hydroxyl [31], and sulfonic acid groups [32] is also effective at enhancing the catalytic ability of the catalysts. Previously, we have reported the electrocatalytic HER activity of metal-corrole complexes bearing pentafluorophenyl [24], nitro [29], hydroxyl [31], and ethoxycarbonyl [33] substituents. The imidazole moiety, as a group with both electron-withdrawing as well as strong proton transfer ability [34,35], may improve the electrocatalytic HER activity of metal–corrole catalysts. In this work, we have prepared a new corrole bearing 4-(1H-imidazole) phenyl group at the 10- position and two strong electron-withdrawing pentafluorophenyl groups at the 5- and 15- positions and its Co, Cu, and Fe complexes (Scheme 1). The results show that the three new metal corroles have good electrocatalytic ability for proton reduction in both organic and aqueous media, with cobalt corrole having the best performance.

2. Results and Discussion

2.1. Synthesis and Characterization

The freebase corrole 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole) phenyl]-corrole(PFIC) was synthesized by modifying the previous method [36]. Metal-corrole complexes were obtained by reacting freebase corrole with different metal salts under certain conditions. All four corroles were synthesized and characterized by ultraviolet-visible spectroscopy (UV-vis), high-resolution mass spectroscopy (HRMS), X-ray photoelectron spectroscopy (XPS), and nuclear magnetic resonance (NMR). All characterization data are listed in Section 3 and the figures are listed in the ESI (Figures S1–S14).

2.1.1. UV-Vis of Freebase Corrole and Metal Complexes

The UV-vis absorption spectra of the freebase corrole and metal complexes are shown in Figure S1. Like the previously reported freebase corrole compound [37], the PFIC has a strong, sharp characteristic absorption peak at about 410 nm (Soret band) and two weaker broad absorption peaks at about 550–650 nm (Q band). PFIC-Co, PFIC-Cu, and PFIC-Fe exhibit the normal Soret and Q absorption bands of metal-corrole complexes. PFIC-Co exhibits a split Soret band and a red-shifted Q band; this is caused by the axial binding of the triphenylphosphine ligand. Such changes can prove that the metal has successfully coordinated with freebase corrole to form metal complexes [38].

2.1.2. Structural Characterization by Single-Crystal X-ray Diffraction

Single-crystal X-rays can clearly reveal the molecular structure of corrole complexes. The single-crystal structure of PFIC-Cu illustrates that it contains an imidazole moiety and exhibits that the Cu ion has successfully coordinated with the four N atoms at the corrole’s center. Also, the Cu ion is in the same plane as the corrole macrocycle (Figure 1). Specific crystal data are detailed in Table S1. The crystal data accurately demonstrate that the lengths of these four Cu-N bonds range from 1.87 to 1.89 Å and that they have an average length of 1.883 Å, which happens to be in agreement with the electronic structure of d8 CuIII [39]. Otherwise, the large angles of ∠N1-Cu-N3 and ∠N2-Cu-N4 produced by the coordinating atoms are 166.0° and 167.7°, respectively, indicating that the Cu is slightly off-center. The PFIC-Cu’s crystallographic information can be obtained for free at the Cambridge Crystallographic Data Centre; its CCDC dentifier is 2301493

2.1.3. X-ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy can effectively detect the elemental distribution of metal complexes. Figure 2a exhibits the full spectrum of PFIC-Co, showing the corresponding signals of Co, N, F, P, and O. In Figure 2b, the peaks of Co 2p3/2 and Co 2p1/2 are at 780.7 and 795.6 eV [40], which exactly correspond to the CoIII oxidation state. Due to the existence of imidazole groups, N can be classified into four categories based on the spectra of N 1s: Graphitic-N, Pyrrolic-N, Co-N, and Pyridinic-N. The corresponding peaks are at 401.1, 400.2, 399.3, and 398.5 eV, respectively [41]. Similarly, the peaks of each element in the spectra of PFIC-Cu and PFIC-Fe are assigned and shown in Figures S13 and S14.

2.2. Cyclic Voltammetry Studies

In the test, all cyclic voltammetry (CV) tests were carried out in N, N-dimethylformamide (DMF) with 0.1 M TBAP as the supporting electrolyte under N2 atmosphere at a scan rate of 100 mV s−1 and a catalyst concentration of 1.0 mM; the working electrode was 3.00 mm glassy carbon, while the counter electrode was a graphite rod, with Ag/AgNO3 serving as the reference electrode. The results are presented in Figure 3. PFIC-Co exhibited an irreversible redox peak and a reversible redox peak in the scanning range from 0 to 2.5 V vs. ferrocene. The first reduction peak, attributed to the reduction of CoIII to CoII, took place at −0.84 V. This phenomenon was irreversible, since the -PPh3 (triphenylphosphine) axial ligand was lost during the reduction process [42]. For PFIC-Cu, two redox peaks appeared at −0.25 and −2.21 V (E1/2), which are attributed to CuIII/CuII and CuII/CuI. Similarly, for PFIC-Fe, its FeIII/FeII and FeII/FeI reduction potentials were −0.97 and −2.14 V (E1/2). The redox peak potentials of all the compounds are summarized in Table 1. A comparison of the MII/MI redox potentials of the three substances shows that PFIC-Co has a more positive potential compared to PFIC-Cu and PFIC-Fe. Furthermore, we also tested the CVs of metal corroles with different scans. The results indicate a positive linear relationship between the redox peak current values of metal corroles and the square root of the scanning speed (Figure S15), implying the electron process to be diffusion-controlled [43].

2.3. Electrocatalytic Activity of Metal Corroles in DMF

In general, the catalytic pathway and the catalytic performance of the catalyst for the HER are closely related to the proton source. The HER catalytic activities of these three metal corroles were tested using acetic acid (AcOH, pKa 13.2 in DMF [44]), trifluoroacetic acid (TFA, pKa 3.5 in DMF [45]), and p-toluene methanesulfonate (TsOH, pKa 2.6 in DMF [44]) as the proton source in DMF.
Firstly, CV tests were performed in an environment where AcOH was the proton source. As shown in Figure 4, with increasing acid equivalents, the CVs of the three complexes gradually became irreversible and the peak currents gradually increased, indicating the presence of a proton reduction process. This was indicated by the gradual disappearance of the second oxidation peaks and the increase in the current of the second reduction peaks. We suggest that the irreversibility is due to the formation of M-H intermediate species in the catalyzing HER, which makes the CV irreversible. It can be seen that the first reduction peak remains almost invariant for all metal corroles, indicating that this step is only the metal reduction process without the involvement of proton reduction reaction. Furthermore, the peak current near the MII/MI increases with the increase in AcOH concentration and reaches a maximum when the acid concentration reaches 32 eq. The position of the reduction also shifts negatively, indicating the involvement of M(I) species in the HER [31].
Figure 5 shows the CVs of three metal corroles in catalyzing proton reduction when using TFA as the proton source. Compared to the AcOH proton source, there is a notably stronger catalytic peak current (−2.5 V) at the equivalent acid concentrations. With an increase in acid concentration, there are no changes observed in the first reduction peaks, while the second redox peak (MII/MI) progressively loses its reversibility. This also implies that M(I) serves as the active center of catalytic proton reduction. It can be found that the catalytic current of PFIC-Fe is the largest, followed by PFIC-Cu and finally PFIC-Co. However, the maximum catalytic current alone is not a comprehensive measure of catalytic performance, and icat/ip values can be introduced to measure catalytic performance (The icat is the peak catalytic current in presence of acid and the ip is the peak current without an acid proton source. The icat/ip ratio may reflect the catalytic activity of the catalysts).
The icat/ip is calculated and presented in Figure 5d; it can be seen that icat increases almost linearly with the increasing acid concentration. PFIC-Co has the largest value of icat/ip, followed by PFIC-Fe and PFIC-Cu. Overpotential is a parameter that can be used to characterize the catalytic property. The overpotential (η) is defined by Equations (1) and (2) [44]; PFIC-Cu exhibits the lowest overpotential of 715 mV, followed by PFIC-Co (813 mV) and PFIC-Fe (898 mV). It can be observed that the catalyst catalyzes the HER differently at different acid concentrations. In the case of PFIC-Co, at low TFA concentration, a reduction peak appeared immediately after the second reduction peak (CoII/CoI), indicating that CoI was further protonated to form CoIII -H. CoIII -H was then additionally reduced and protonated to form CoII. When the concentration of TFA increased to a certain value, the overpotential shifted to the positive direction and only one reduction peak appeared, indicating that, at this time, CoII was the active center of proton reduction, and followed by the direct formation of CoIV-H species via the protonation of CoII [28], which could then be reduced to CoIII-H and finally complete the hydrogen evolution via the Heyrovsky process.
η = E H A 0 E
E H A 0 = E H + 0 2.303 R T F p K a H A
E H A 0 is the standard electrode potential of H+/H2, E H + 0 is the standard potential for the solvated proton–dihydrogen couple E H + 0 = −0.62 V in DMF, p K a H A is the dissociation constant of acid in solvent, and E T F A 0 = −0.83 V is calculated using Equation (2) [33].
From Figure 6, it can be seen that when using TsOH as the proton source, we can find that the catalytic process of the three catalysts is the same as TFA. This also indicates that the catalytic pathways of the catalysts are the same in these two proton sources. In comparison with using TFA as the proton source, the peak current and the icat/ip become larger for the same acid equivalent, which is due to the smaller pKa value of TsOH.
The CVs of bare GC electrode without catalysts and in the presence of catalysts under the same acid concentration is shown in Figure S16. The results clearly show that the sharp catalytic current increasing is indeed a result of the electrocatalysis of the molecular catalysts. The catalytic efficiency (C.E) and turnover frequency (TOF) were calculated by using the following Equations (3) and (5), respectively [45]:
C . E = i c a t / i p C H A / C c a t
k o b s = k [ H + ] x
k o b s = 1.94 v ( i c a t i p ) 2
where icat is the peak current in acid and ip is the peak current in non-acid. CHA denotes the concentration of organic acid, Ccat refers to the concentration of the catalyst, ν denotes the scanning rate, and kobs represents the rate constant or turnover frequency (TOF) of the catalytic reaction. The TOFs of PFIC-Co, PFIC-Cu, and PFIC-Fe in TsOH (32 equiv.) are calculated to be 265 s−1, 138 s−1, and 109 s−1, respectively. PFIC-Co has the largest TOF, indicating that it has the best catalytic activity. Other parameters of the three corroles at TFA and TsOH are summarized in Tables S2 and S3. To compare the HER activity of PFIC-Co, PFIC-Cu, and PFIC-Fe with other molecular electrocatalysts, the previously reported HER activities of some transition metal molecular electrocatalysts are summarized in Table S4 [28,31,46,47,48,49,50]. The table shows that the performance of these molecular catalysts is diverse. The overpotentials of these catalysts are generally around 800 mV, which is at the same level as that of our catalysts. The TOF, on the other hand, varies from high to low, with an average level of around 150 s−1, where the highest can reach up to 1350 s−1 and the lowest is only 13 s−1. It can be seen that these three metal-corrole catalysts show good catalytic performance. However, there is still a gap compared to the state-of-the-art molecular electrocatalysts.

2.4. Possible Catalytic Pathways for Hydrogen Production

Electrochemical hydrogen production pathways are generally inseparable from proton transfer and electron transfer processes, and most of the active center is metal [51]. Based on previous reports on electrocatalytic hydrogen production by metal corroles [47,48], the speculated catalytic pathways of the current system are depicted in Scheme 2. For PFIC-Co, when the proton source is the weak acid AcOH, CoII/CoI gradually becomes irreversible with the increase in acid concentration, suggesting that CoI is involved in the HER. The formed CoI-H may rapidly convert to CoIII-H, followed by subsequent reduction to form CoII-H at a more negative potential. The final protonation of CoII-H leads to hydrogen release and finishes the catalytic cycle, so the possible catalytic pathway is EECEC (E: electron transfer, C: proton transfer, I-II-III-IV-V, in Scheme 2). When using low concentrations of TFA or TsOH as a proton source, the possible catalytic pathway for hydrogen production is also EECEC. However, at high concentrations, the active center changes to CoII, which is first protonated to CoIV-H, and then CoIV-H is subsequently reduced to CoIII-H. Finally, CoIII-H acquires a proton to produce H2, so the possible catalytic pathway is ECEC (I-VI-VII-VIII). For PFIC-Fe, the possible catalytic pathways are consistent with PFIC-Co. However, as for PFIC-Cu, at low acid concentrations in all acids, the active center is CuI, so the possible catalytic pathway is EECEC. Whereas at high concentrations, the active center is CuII and the possible catalytic pathway is ECEC.

2.5. Electrocatalytic Proton Reduction in Neutral Aqueous Media

The catalytic properties and catalytic pathways of metal corroles in organic systems have been explored above. From the point of view of practical application, further investigation of the catalytic process of catalysts for HER in aqueous solutions is necessary. We carried out electrochemical studies using three metal corroles as catalysts in a buffer solution of pH = 7 containing 0.25 M phosphate. In this system, Ag/AgCl was utilized as the reference electrode, while the remaining electrode systems corresponded to that of the organic phase system. Due to the low solubility of organic molecules in pure water, a co-solution of acetonitrile and water was used (Vwater/Vacetonitrile = 3/2). As shown in Figure 7, a negligible current was produced in the absence of any catalyst, with the overpotential at approximately −1.14 V versus the reversible hydrogen electrode (RHE). After the addition of 5.0 μm/L of catalysts to the buffer solution (10 mL), it was found that the catalytic current increased significantly and the overpotential shifted positively to some extent. Among them, the catalytic current of PFIC-Co was the highest (0.91 mA) and the overpotential was positively shifted by 380 mV. The results suggest that the three catalysts exhibit effective catalytic activity in a neutral aqueous solution. In addition, Figure S17 demonstrates the effect of different catalyst concentrations on catalytic performance, and it was found that the catalytic current gradually increased with the increase in catalyst dosage. And we also conducted controlled potential electrolysis (CPE) experiments for the three catalysts at different overpotentials (738 mV–1138 mV), according to Equation (6) [52]. From Figure S18, it can be found that the TOF increases as the overpotential increases. When the overpotential reaches 1138 mV, the TOFs of PFIC-Co, PFIC-Cu, and PFIC-Fe are 405, 310, and 272 h−1, respectively. Also, the Faraday efficiency of electrocatalytic hydrogen evolution could be calculated using Equations (7) and (8) [53,54]. It turned out that the Faraday efficiencies of PFIC-Fe, PFIC-Cu, and PFIC-Co were 89.8%, 91.3%, and 94.7%, respectively. The Faraday efficiencies of some transition metal complexes reported previously are summarized in Table S5 [55,56]. The table shows that most of the complexes have a Faraday efficiency of up to 85% or more, with the lowest being 74% and the highest being up to 98%. This shows that these three catalysts have good hydrogen production ability. In addition, the stability of the three catalysts was tested using chronoamperometry (Figure S19). These metal corroles exhibited good stability over long periods of electrolysis.
T O F = Δ C F × n 1 × n 2 × t
p H t h e o r e t i c a l = 14 + l g I t F V
F E ( % ) = A V G ( p H p r a c t i c a l p H t h e o r e t i c a l )
where ΔC represents the increase in charge compared to the blank (without catalyst), with F representing the Faraday constant, n1 being the number of electrons required for the production of 1 mol of H2 (n1 = 2 in this electrolysis), n2 indicating the mol of catalyst used, and t representing the electrolysis time. p H t h e o r e t i c a l stands for theoretical pH value, I is the current intensity, V is the electrolyte volume, and p H p r a c t i c a l represents the practical pH value.

3. Materials and Methods

All reagents were purchased commercially, are of high purity, and can be used without further purification. UV-visible spectra were measured using a Hitachi U-3010. High-resolution mass spectrometry was performed with a Bruker maxis impact mass spectrometer that utilized an electrospray (ESI) source. The Bruker Advance III 400 MHz spectrometer was utilized to detect 1H-NMR, 31P NMR, and 19F NMR spectra in CDCl3 (d = 7.26 ppm) solvent at room temperature. Additionally, X-ray photoelectron spectroscopy (XPS) was recorded using an Axis Ultra-DLD spectrophotometer. Single Crystal X-ray Diffraction was measured using the Rigaku XtalAB PRO MM007DW. The CVs were measured with a CHI-660E electro-chemical analyzer in a DMF system and TBAP as a supporting electrolyte under N2 protection for 30 min. The CPE experiments took place in a two-room electrolytic cell filled with a blend of acetonitrile and water (2:3 ratio) mixed with 0.1 M KCl as the electrolyte. Ag/AgCl electrodes served as the reference electrodes, with a GC plate (10 × 10 mm) used as the working electrode in one room, while a platinum wire was placed in another room.
  • Synthesis of 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole
Freebase corrole was synthesized using an improved method published in the literature. First, 160 mL of freshly steamed pyrrole and 2.8 g of pentafluorobenzaldehyde were mixed homogeneously, followed by the addition of 120 μL of trifluoroacetic acid (TFA) as a catalyst, and the reaction was carried out for 45 min at room temperature. Next, 240 μL of ethylenediamine was added to stop the reaction. The extra pyrrole was removed through distillation under reduced pressure at 140 °C, resulting in an oily product. To obtain a purer product, the crude product was separated via column chromatography with eluents of CH2Cl2 and Hex (VDCM/VHex = 3:1). The dipyrromethane was obtained at a yield of 83%. Next, dipyrromethane (1.25 g, 4 mmol) and 4-(1H-Imidazol-1-yl) benzaldehyde (344 mg, 2 mmol) were dissolved in 300 mL of DCM, followed by the addition of 460 μL of trifluoroacetic acid, and the reaction was carried out for 4 h at an ambient temperature protected from light. Immediately thereafter, 920 μL of triethylamine and DDQ (1.18 g, 4 mmol) was added. After 1 h of reaction, the crude product was obtained by removing the DCM using a rotary evaporator, followed by separating the resulting product through a first-column chromatography (silica gel: 100 mesh, eluent: DCM) to eliminate the DDQ. Then, the product was once again isolated using column chromatography with silica gel (400 mesh) and an eluent of dichloromethane and ethyl acetate in a ratio of 3:1, and the reddish-orange portion of the product solution was collected and the solvent was removed by rotatable evaporation under vacumn to give purple product (200 mg, yield: 11%). 1H NMR (500 MHz, Chloroform-d) δ 9.13 (s, 2H), 8.78–8.64 (m, 4H), 8.58 (s, 2H), 8.34 (d, J = 8.6 Hz, 2H), 7.95 (s, 1H), 7.82 (d, J = 7.8 Hz, 2H), 7.61 (s, 1H), 7.38 (s, 1H). 19F NMR (471 MHz, Chloroform-d) δ −137.90 (4F), −152.58 (2F), −161.67 (4F). HRMS (ESI) [M + H]+, calculated for C40H19F10N6: 773.1506, found: 773.1520.
  • Synthesis of cobalt 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole
Freebase corrole (50 mg, 65 μmol) and NaOAc (60 mg, 732 μmol) were dissolved in 30 mL of ethanol and stirred for 5 min at room temperature, followed by the addition of Co(OAc)2·4H2O (50 mg, 200 μmol) and triphenylphosphine (100 mg, 382 μmol), and the reaction was continued for two hours. Then, 30 mL of dichloromethane and 30 mL of water were added to extract, and the organic phase was collected and dried using sodium sulfate. Column chromatography using CH2Cl2 as the eluent was used to further purify the raw material obtained. The resulting product was recrystallized using CH2Cl2 and Hex for further purification. The clean substance (red-brown powder) was obtained by recrystallizing, resulting in a 75% yield of 53.13 mg. 1H NMR (500 MHz, Chloroform-d) δ 8.69 (d, J = 39.8 Hz, 2H), 8.22 (s, 2H), 8.00 (t, J = 35.0 Hz, 4H), 7.72–7.31 (m, 7H), 7.04–6.88 (m, 3H), 6.57 (d, J = 35.2 Hz, 6H), 4.66–4.36 (m, 6H). 19F NMR (471 MHz, Chloroform-d) δ −136.71–−138.45 (4F), −154.44 (2F), −161.88–−164.84 (4F). HRMS (ESI) [M + H]+, calculated for C58H31CoF10N6P: 1091.1515, found: 1091.1551.
  • Synthesis of copper 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole
Freebase corrole (56 mg, 72 μmol) was dissolved in 50 mL of ethanol, and NaOAc (60 mg, 732 μmol) was added, stirred at room temperature for 10 min and then Cu(OAc)2 -H2O (70 mg, 350 μmol) was added, followed by the reaction for 1.5 h, and the liquid was dried completely to acquire the unfinished product, which was later divided using column chromatography, and CH2Cl2 was used as the eluent to obtain the purer product. A yellow-black powder (45 mg, yield: 75.1%) was obtained. 1H NMR (500 MHz, Chloroform-d) δ 8.05–7.95 (m, 2H), 7.84 (d, J = 40.3 Hz, 4H), 7.47–7.40 (m, 2H), 7.35 (d, J = 2.7 Hz, 1H), 7.23 (s, 2H), 7.17–7.05 (m, 3H), 6.99 (s, 1H). HRMS (ESI) [M + H]+, calculated for C40H16CuF10N6: 833.0567, found: 833.0570.
  • Synthesis of iron 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole
The freebase corrole weighing 50 mg (65 μmol) and ferrous chloride tetrahydrate weighing 2.389 g (13 mmol) were combined with 30 mL DMF and reacted for 6 h at 140 °C in a nitrogen-protected atmosphere. Waiting until the reaction solution had cooled down to room temperature, we extracted it with CH2Cl2 (50 mL) and H2O (50 mL). Then, we collected the organic phase and washed it with saturated brine. Finally, we dried it with Na2SO4. The product was then purified by chromatography in a similar procedure to the preparation of cobalt corrole with CH2Cl2 as the eluent. Finally, a dark brown solid was obtained (52 mg, yield: 82%). HRMS (ESI) [M + H]+, calculated for C40H16F10FeN6: 826.0621, found: 826.0629.

4. Conclusions

In summary, the new 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole containing an imidazole moiety and its cobalt, copper and iron complexes had been synthesized and well characterized. Their reactivity for electrocatalytic proton reduction in different media was also investigated. The three prepared metal corroles showed good catalytic performance in both aqueous and organic phase media. The Co corrole exhibited the best performance with a TOF of up to 265 s−1 in DMF (TsOH proton source). We also explored the catalytic proton reduction in organic solvents with different acids as the proton source for each catalyst. It turned out that the electrocatalytic HER by these metal corroles may proceed via two competing ECEC and/or EECEC pathways depending on the acid proton sources used.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14010005/s1, Figure S1: UV-vis spectrum of PFIC and PFIC-Co, PFIC-Cu and PFIC-Fe in CH2Cl2; Figure S2: 1H NMR spectrum of 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S3: 19F NMR spectrum of 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S4: 1H NMR spectrum of cobalt 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S5: 19F NMR spectrum of cobalt 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S6: 31P NMR spectrum of cobalt 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S7: 1H NMR spectrum of copper 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S8: 19F NMR spectrum of copper 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S9: High resolution mass spectrum of 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S10: High resolution mass spectrum of cobalt 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S11: High resolution mass spectrum of copper 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole) phenyl]-corrole; Figure S12: High resolution mass spectrum of iron 5,15-bis-(pentafluorophenyl)-10-[4-(1H-imidazole)phenyl]-corrole; Figure S13: XPS survey spectrum of PFIC-Cu (a); XPS spectra of Cu 2p (b) and N 1s (c) and F 1s (d) of PFIC-Cu; Figure S14: XPS survey spectrum of PFIC-Fe (a); XPS spectra of Fe 2p (b) and N 1s (c) and F 1s (d) of PFIC-Fe; Figure S15: Plots of metal corrole complexes Co, Cu, and Fe (a–c, 1.0 mM) for sweep rate (v) variations of 100–400 mV/s and peak currents (ip) versus square root of the sweep rate (v1/2) for the reduction and first oxidation peaks (d–f); Figure S16: Comparison of CV test between bare GC and PFIC-Co, PFIC-Cu and PFIC-Fe at high acid concentration; Figure S17: CV plots (0.00–5.00 µM) of different concentrations of metallcorrole complexes (PFIC-Co(a), PFIC-Cu(b), PFIC-Fe(c)) in a mixed system of acetonitrile and water (pH = 7); Figure S18: (a) Charge increase for PFIC-Co (5.00 μM), (b) Charge increase for PFIC-Cu (5.00 μM), (c) Charge increase for PFIC-Fe (5.00 μM) at different overpotentials; Figure S19: Catalytic current versus time obtained from 4 h CPE with 5.00 μM (a) PFIC-Co, (b) PFIC-Cu (c) PFIC-Fe in buffer solution at −1.04 V vs. RHE; Table S1: Crystal data and structure refinement for PFEC-Co; Table S2: Catalytic performance parameters of three metal complexes in TFA system; Table S3: Catalytic performance parameters of three metal complexes in TsOH system; Table S4: HER activity for transition metal corroles in organic solvent by using organic acids as proton; Table S5: Faraday efficiency of transition metal complexes in nature homogeneous aqueous solution. [28,31,52,53,54,55,56,57,58].

Author Contributions

Conceptualization, L.-W.W., Y.-W.R. and H.-Y.L.; visualization, Y.-F.Y., S.-Y.X. and X.-Y.C.; writing—original draft preparation, L.-W.W.; writing—review and editing, H.-Y.L. supervision, H.-Y.L. and L.-P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 22005052, 20161068) and the Research Fund Program of Guangdong Provincial Key Laboratory of Fuel Cell Technology (Grant No. FC202211).

Data Availability Statement

The data that support the findings of this study are available from the Supplementary Materials.

Acknowledgments

We thank the School of Chemistry and Chemical Engineering of South China University of Technology for their generous support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armaroli, N.; Balzani, V. The Hydrogen Issue. ChemSusChem 2011, 4, 21–36. [Google Scholar] [CrossRef] [PubMed]
  2. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef] [PubMed]
  3. Sazali, N. Emerging technologies by hydrogen: A review. Int. J. Hydrogen Energy 2020, 45, 18753–18771. [Google Scholar] [CrossRef]
  4. Hosseini, S.E.; Butler, B. An overview of development and challenges in hydrogen powered vehicles. Int. J. Green Energy 2020, 17, 13–37. [Google Scholar] [CrossRef]
  5. Alshammari, B.H.; Begum, H.; Ibrahim, F.A.; Hamdy, M.S.; Oyshi, T.A.; Khatun, N.; Hasnat, M.A. Electrocatalytic Hydrogen Evolution Reaction from Acetic Acid over Gold Immobilized Glassy Carbon Surface. Catalysts 2023, 13, 744. [Google Scholar] [CrossRef]
  6. Xie, L.; Tian, J.; Ouyang, Y.; Guo, X.; Zhang, W.; Apfel, U.; Cao, R. Water-Soluble Polymers with Appending Porphyrins as Bioinspired Catalysts for the Hydrogen Evolution Reaction. Angew. Chem. Int. Ed. 2020, 59, 15844–15848. [Google Scholar] [CrossRef] [PubMed]
  7. Guo, X.; Wang, N.; Li, X.; Zhang, Z.; Zhao, J.; Ren, W.; Ding, S.; Xu, G.; Li, J.; Apfel, U.; et al. Homolytic versus Heterolytic Hydrogen Evolution Reaction Steered by a Steric Effect. Angew. Chem. Int. Ed. 2020, 59, 8941–8946. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, X.; An, X.; Tang, L.; Chen, T.; Zhang, G. Confining platinum clusters in ZIF-8-derived porous N-doped carbon arrays for high-performance hydrogen evolution reaction. Chem. Eng. J. 2022, 429, 132259. [Google Scholar] [CrossRef]
  9. Xin, Y.; Wang, F.; Chen, L.; Li, Y.; Shen, K. Superior bifunctional cobalt/nitrogen-codoped carbon nanosheet arrays on copper foam enable stable energy-saving hydrogen production accompanied with glucose upgrading. Green Chem. 2022, 24, 6544–6555. [Google Scholar] [CrossRef]
  10. Li, H.; Li, X.; Lei, H.; Zhou, G.; Zhang, W.; Cao, R. Convenient Immobilization of Cobalt Corroles on Carbon Nanotubes through Covalent Bonds for Electrocatalytic Hydrogen and Oxygen Evolution Reactions. Chemsuschem 2019, 12, 801–806. [Google Scholar] [CrossRef]
  11. Lemon, C.M.; Huynh, M.; Maher, A.G.; Anderson, B.L.; Bloch, E.D.; Powers, D.C.; Nocera, D.G. Electronic Structure of Copper Corroles. Angew. Chem.-Int. Ed. 2016, 55, 2176–2180. [Google Scholar] [CrossRef] [PubMed]
  12. Peng, X.Y.; Han, J.X.; Li, X.L.; Liu, G.J.; Xu, Y.H.; Peng, Y.X.; Nie, S.; Li, W.Z.; Li, X.R.; Chen, Z.; et al. Electrocatalytic hydrogen evolution with a copper porphyrin bearing meso-(o-carborane) substituents. Chem. Commun. 2023, 59, 10777–10780. [Google Scholar] [CrossRef] [PubMed]
  13. Islam, M.N.; Ahmed, J.; Faisal, M.; Algethami, J.S.; Aoki, K.; Nagao, Y.; Harraz, F.A.; Hasnat, M.A. Efficient Electrocatalytic Hydrogen Evolution Reaction on CuO Immobilized Stainless-Steel Electrode Prepared by the SILAR Method. ChemistrySelect 2023, 8, e202301077. [Google Scholar] [CrossRef]
  14. Tatematsu, R.; Inomata, T.; Ozawa, T.; Masuda, H. Electrocatalytic Hydrogen Production by a Nickel(II) Complex with a Phosphinopyridyl Ligand. Angew. Chem. Int. Ed. 2016, 55, 5247–5250. [Google Scholar] [CrossRef] [PubMed]
  15. Simkhovich, L.; Mahammed, A.; Goldberg, I.; Gross, Z. Synthesis and Characterization of Germanium, Tin, Phosphorus, Iron, and Rhodium Complexes of Tris(pentafluorophenyl)corrole, and the Utilization of the Iron and Rhodium Corroles as Cyclopropanation Catalysts. Chem.—A Eur. J. 2001, 7, 1041–1055. [Google Scholar] [CrossRef]
  16. Zhang, W.; Lai, W.Z.; Cao, R. Energy-Related Small Molecule Activation Reactions: Oxygen Reduction and Hydrogen and Oxygen Evolution Reactions Catalyzed by Porphyrin- and Corrole-Based Systems. Chem. Rev. 2017, 117, 3717–3797. [Google Scholar] [CrossRef] [PubMed]
  17. Joseph, M.; Haridas, S. Recent progresses in porphyrin assisted hydrogen evolution. Int. J. Hydrogen Energy 2020, 45, 11954–11975. [Google Scholar] [CrossRef]
  18. Yu, B.; Chen, X.; Zhao, Y.; Chen, W.; Xiao, X.; Liu, H. Graphene Oxide-based Cobalt Porphyrin Composites for Electrocatalytic Hydrogen Evolution Reaction. Chem. J. Chin. Univ.-Chin. 2022, 43, 20210549. [Google Scholar]
  19. Beyene, B.B.; Hung, C.-H. Recent progress on metalloporphyrin-based hydrogen evolution catalysis. Coord. Chem. Rev. 2020, 410, 213234. [Google Scholar] [CrossRef]
  20. Li, X.; Lei, H.; Xie, L.; Wang, N.; Zhang, W.; Cao, R. Metalloporphyrins as Catalytic Models for Studying Hydrogen and Oxygen Evolution and Oxygen Reduction Reactions. Acc. Chem. Res. 2022, 55, 878–892. [Google Scholar] [CrossRef]
  21. Qi, X.W.; Yang, G.; Guo, X.S.; Si, L.P.; Zhang, H.; Liu, H.Y. Electrocatalytic Hydrogen Evolution by Water-Soluble Cobalt (II), Copper (II) and Iron (III) meso-Tetrakis(carboxyl)porphyrin. Eur. J. Inorg. Chem. 2023, 26, 8. [Google Scholar] [CrossRef]
  22. Di Natale, C.; Gros, C.P.; Paolesse, R. Corroles at work: A small macrocycle for great applications. Chem. Soc. Rev. 2022, 51, 1277–1335. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, D.-X.; Yuan, H.-Q.; Wang, H.-H.; Ali, A.; Wen, W.-H.; Xie, A.-N.; Zhan, S.-Z.; Liu, H.-Y. Transition metal tetrapentafluorophenyl porphyrin catalyzed hydrogen evolution from acetic acid and water. Transit. Met. Chem. 2017, 42, 773–782. [Google Scholar] [CrossRef]
  24. Wan, B.; Cheng, F.; Lan, J.; Zhao, Y.; Yang, G.; Sun, Y.M.; Si, L.P.; Liu, H.Y. Electrocatalytic hydrogen evolution of manganese corrole. Int. J. Hydrogen Energy 2023, 48, 5506–5517. [Google Scholar] [CrossRef]
  25. Lv, H.; Zhang, X.-P.; Guo, K.; Han, J.; Guo, H.; Lei, H.; Li, X.; Zhang, W.; Apfel, U.-P.; Cao, R. Coordination Tuning of Metal Porphyrins for Improved Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2023, 62, e202305938. [Google Scholar] [CrossRef] [PubMed]
  26. Lei, H.T.; Li, X.L.; Meng, J.; Zheng, H.Q.; Zhang, W.; Cao, R. Structure Effects of Metal Corroles on Energy-Related Small Molecule Activation Reactions. Acs Catal. 2019, 9, 4320–4344. [Google Scholar] [CrossRef]
  27. Mahammed, A.; Mondal, B.; Rana, A.; Dey, A.; Gross, Z. The cobalt corrole catalyzed hydrogen evolution reaction: Surprising electronic effects and characterization of key reaction intermediates. Chem. Commun. 2014, 50, 2725–2727. [Google Scholar] [CrossRef]
  28. Liu, Z.Y.; Lai, J.W.; Yang, G.; Ren, B.P.; Lv, Z.Y.; Si, L.P.; Zhang, H.; Liu, H.Y. Electrocatalytic hydrogen production by CN-substituted cobalt triaryl corroles. Catal. Sci. Technol. 2022, 12, 5125–5135. [Google Scholar] [CrossRef]
  29. Chen, H.; Huang, D.L.; Hossain, M.S.; Luo, G.T.; Liu, H.Y. Electrocatalytic activity of cobalt tris(4-nitrophenyl)corrole for hydrogen evolution from water. J. Coord. Chem. 2019, 72, 2791–2803. [Google Scholar] [CrossRef]
  30. Beyene, B.B.; Mane, S.B.; Leonardus, M.; Hung, C.H. Effects of Position and Electronic Nature of Substituents on Cobalt-Porphyrin-Catalyzed Hydrogen Evolution Reaction. ChemistrySelect 2017, 2, 10565–10571. [Google Scholar] [CrossRef]
  31. Lv, Z.-Y.; Yang, G.; Ren, B.-P.; Liu, Z.-Y.; Zhang, H.; Si, L.-P.; Liu, H.-Y.; Chang, C.-K. Electrocatalytic Hydrogen Evolution of the Cobalt Triaryl Corroles Bearing Hydroxyl Groups. Eur. J. Inorg. Chem. 2023, 26, e202200755. [Google Scholar] [CrossRef]
  32. Wang, N.; Zhang, X.P.; Han, J.X.; Lei, H.T.; Zhang, Q.X.; Zhang, H.; Zhang, W.; Apfel, U.P.; Cao, R. Promoting hydrogen evolution reaction with a sulfonic proton relay. Chin. J. Catal. 2023, 45, 88–94. [Google Scholar] [CrossRef]
  33. Peng, W.Y.; Lan, J.; Zhu, Z.M.; Si, L.P.; Zhang, H.; Zhan, S.Z.; Liu, H.Y. Synthesis of metal (Ga, Co and Fe) 5,15-bis(pentafluorophenyl)-10-ethox-ycarbonylcorrole and their electrocatalytic hydrogen evolution activity. Inorg. Chem. Commun. 2022, 140, 109453. [Google Scholar] [CrossRef]
  34. Sharma, V.K.; Mahammed, A.; Mizrahi, A.; Morales, M.; Fridman, N.; Gray, H.B.; Gross, Z. Dimeric Corrole Analogs of Chlorophyll Special Pairs. J. Am. Chem. Soc. 2021, 143, 9450–9460. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, J.D.; Li, P.; Li, X.L.; Xie, L.S.; Wang, N.; Lei, H.T.; Zhang, C.C.; Zhang, W.; Lee, Y.M.; Zhang, W.Q.; et al. Crucial Roles of a Pendant Imidazole Ligand of a Cobalt Porphyrin Complex in the Stoichiometric and Catalytic Reduction of Dioxygen. Angew. Chem. Int. Ed. 2022, 61, 9. [Google Scholar]
  36. Koszarna, B.; Gryko, D.T. Efficient synthesis of meso-substituted corroles in a H2O-MeOH mixture. J. Org. Chem. 2006, 71, 3707–3717. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, G.; Ullah, Z.; Yang, W.; Kwon, H.W.; Liang, Z.X.; Zhan, X.; Yuan, G.Q.; Liu, H.Y. Substituent Effect on Ligand-Centered Electrocatalytic Hydrogen Evolution of Phosphorus Corroles. ChemSusChem 2023, 16, e202300211. [Google Scholar] [CrossRef] [PubMed]
  38. Liang, Y.-Y.; Li, M.-Y.; Shi, L.; Lin, D.-Z.; Zhan, S.-Z.; Liu, H.-Y. Electrocatalytic hydrogen evolution by cobalt triaryl corroles with appended ester and carboxyl on the 10-phenyl group. J. Coord. Chem. 2021, 74, 1414–1424. [Google Scholar] [CrossRef]
  39. Lei, H.T.; Fang, H.Y.; Han, Y.Z.; Lai, W.Z.; Fu, X.F.; Cao, R. Reactivity and Pathway Studies of Hydrogen Evolution Catalyzed by Copper Corroles. ACS Catal. 2015, 5, 5145–5153. [Google Scholar] [CrossRef]
  40. Ding, D.N.; Shen, K.; Chen, X.D.; Chen, H.R.; Chen, J.Y.; Fan, T.; Wu, R.F.; Li, Y.W. Multi-Level Architecture Optimization of MOF-Templated Co-Based Nanoparticles Embedded in Hollow N-Doped Carbon Polyhedra for Efficient OER and ORR. ACS Catal. 2018, 8, 7879–7888. [Google Scholar] [CrossRef]
  41. Wang, Z.Q.; Shen, K.; Chen, L.Y.; Li, Y.W. Scalable synthesis of multi-shelled hollow N-doped carbon nanosheet arrays with confined Co/CoP heterostructures from MOFs for pH-universal hydrogen evolution reaction. Sci. China-Chem. 2022, 65, 619–629. [Google Scholar] [CrossRef]
  42. Li, B.; Ou, Z.; Meng, D.; Tang, J.; Fang, Y.; Liu, R.; Kadish, K.M. Cobalt triarylcorroles containing one, two or three nitro groups. Effect of NO2 substitution on electrochemical properties and catalytic activity for reduction of molecular oxygen in acid media. J. Inorg. Biochem. 2014, 136, 130–139. [Google Scholar] [CrossRef] [PubMed]
  43. Thoi, V.S.; Karunadasa, H.I.; Surendranath, Y.; Long, J.R.; Chang, C.J. Electrochemical generation of hydrogen from acetic acid using a molecular molybdenum-oxo catalyst. Energy Environ. Sci. 2012, 5, 7762–7770. [Google Scholar] [CrossRef]
  44. DiRisio, R.J.; Armstrong, J.E.; Frank, M.A.; Lake, W.R.; McNamara, W.R. Cobalt Schiff-base complexes for electrocatalytic hydrogen generation. Dalton Trans. 2017, 46, 10418–10425. [Google Scholar] [CrossRef] [PubMed]
  45. Stewart, M.P.; Ho, M.H.; Wiese, S.; Lindstrom, M.L.; Thogerson, C.E.; Raugei, S.; Bullock, R.M.; Helm, M.L. High Catalytic Rates for Hydrogen Production Using Nickel Electrocatalysts with Seven-Membered Cyclic Diphosphine Ligands Containing One Pendant Amine. J. Am. Chem. Soc. 2013, 135, 6033–6046. [Google Scholar] [CrossRef] [PubMed]
  46. Beyene, B.B.; Hung, C.H. Porphyrin-Based Electrochemical H2 Evolution: Role of Central Metal Ion on Overpotential and Catalytic Activity. Electrocatalysis 2018, 9, 689–696. [Google Scholar] [CrossRef]
  47. Kumar, A.; Fite, S.; Raslin, A.; Kumar, S.; Mizrahi, A.; Mahammed, A.; Gross, Z. Beneficial Effects on the Cobalt-Catalyzed Hydrogen Evolution Reaction Induced by Corrole Chelation. ACS Catal. 2023, 13, 13344–13353. [Google Scholar] [CrossRef]
  48. Chen, Q.-C.; Fite, S.; Fridman, N.; Tumanskii, B.; Mahammed, A.; Gross, Z. Hydrogen Evolution Catalyzed by Corrole-Chelated Nickel Complexes, Characterized in all Catalysis-Relevant Oxidation States. ACS Catal. 2022, 12, 4310–4317. [Google Scholar] [CrossRef]
  49. Karunadasa, H.I.; Montalvo, E.; Sun, Y.; Majda, M.; Long, J.R.; Chang, C.J. A Molecular MoS2 Edge Site Mimic for Catalytic Hydrogen Generation. Science 2012, 335, 698–702. [Google Scholar] [CrossRef]
  50. Drosou, M.; Kamatsos, F.; Mitsopoulou, C.A. Recent advances in the pathways of the hydrogen evolution reaction by non-innocent sulfur-coordinating metal complexes. Inorg. Chem. Front. 2020, 7, 37–71. [Google Scholar] [CrossRef]
  51. Costentin, C.; Drouet, S.; Robert, M.; Savéant, J.-M. Turnover Numbers, Turnover Frequencies, and Overpotential in Molecular Catalysis of Electrochemical Reactions. Cyclic Voltammetry and Preparative-Scale Electrolysis. J. Am. Chem. Soc. 2012, 134, 11235–11242. [Google Scholar] [CrossRef] [PubMed]
  52. Sudhakar, K.; Mahammed, A.; Chen, Q.-C.; Fridman, N.; Tumanskii, B.; Gross, Z. Copper Complexes of CF3-Substituted Corroles for Affecting Redox Potentials and Electrocatalysis. ACS Appl. Energy Mater. 2020, 3, 2828–2836. [Google Scholar] [CrossRef]
  53. Yadav, P.; Nigel-Etinger, I.; Kumar, A.; Mizrahi, A.; Mahammed, A.; Fridman, N.; Lipstman, S.; Goldberg, I.; Gross, Z. Hydrogen evolution catalysis by terminal molybdenum-oxo complexes. iScience 2021, 24, 102924. [Google Scholar] [CrossRef] [PubMed]
  54. Dolganov, A.V.; Chernyaeva, O.; Kostryukov, S.; Balandina, A.; Solovyova, E.; Yudina, A.; Akhmatova, A.; Lyukshina, Y. Control of the substituent at the nitrogen atom in a 2,4,6-triphenylpyridinium perchlorates tunes the electrocatalytic hydrogen evolution mechanism and efficiency. Int. J. Hydrogen Energy 2020, 45, 501–507. [Google Scholar] [CrossRef]
  55. Dolui, D.; Khandelwal, S.; Shaik, A.; Gaat, D. Enzyme-Inspired Synthetic Proton Relays Generate Fast and Acid-Stable Cobalt-Based H2 Production Electrocatalysts. ACS Catal. 2019, 9, 10115–10125. [Google Scholar] [CrossRef]
  56. Dolui, D.; Mir, A.Q.; Dutta, A. Probing the peripheral role of amines in photo- and electrocatalytic H2 production by molecular cobalt complexes. Chem. Commun. 2020, 56, 14841–14844. [Google Scholar] [CrossRef]
  57. Zhong, Y.-Q.; Hossain, S.; Chen, Y.; Fan, Q.-H.; Zhan, S.-Z.; Liu, H.-Y. A comparative study of electrocatalytic hydrogen evolution by iron complexes of corrole and porphyrin from acetic acid and water. Transit. Met. Chem. 2019, 44, 399–406. [Google Scholar] [CrossRef]
  58. Wang, C.-L.; Yang, H.; Du, J.; Zhan, S.-Z. A new catalyst based on a nickel(II) complex of diiminodiphosphine for hydrogen evolution and oxidation. Int. J. Hydrogen Energy 2021, 46, 32480–32489. [Google Scholar] [CrossRef]
Scheme 1. The structure of the investigated metal corroles.
Scheme 1. The structure of the investigated metal corroles.
Catalysts 14 00005 sch001
Figure 1. Thermal ellipsoid plots (50% probability) of the X-ray structure of PFIC-Cu.
Figure 1. Thermal ellipsoid plots (50% probability) of the X-ray structure of PFIC-Cu.
Catalysts 14 00005 g001
Figure 2. XPS survey spectrum of PFIC-Co (a); XPS spectra of Co 2p (b), N 1s (c), and P 2p (d).
Figure 2. XPS survey spectrum of PFIC-Co (a); XPS spectra of Co 2p (b), N 1s (c), and P 2p (d).
Catalysts 14 00005 g002
Figure 3. CVs of the metal complexes PFIC-Co, PFIC-Cu, and PFIC-Fe (1.0 mM) in DMF (0.1 M TBAP).
Figure 3. CVs of the metal complexes PFIC-Co, PFIC-Cu, and PFIC-Fe (1.0 mM) in DMF (0.1 M TBAP).
Catalysts 14 00005 g003
Figure 4. CV curves of compounds PFIC-Co (a), PFIC-Cu (b), and PFIC-Fe (c) (1.00 mM) with increasing AcOH content (0–32 equiv.).
Figure 4. CV curves of compounds PFIC-Co (a), PFIC-Cu (b), and PFIC-Fe (c) (1.00 mM) with increasing AcOH content (0–32 equiv.).
Catalysts 14 00005 g004
Figure 5. CVs of metal corroles (ac), 1.0 mM) in different concentrations of TFA (0–32 eq.), and variation of icat/ip values in DMF solutions containing different concentrations of TFA (d).
Figure 5. CVs of metal corroles (ac), 1.0 mM) in different concentrations of TFA (0–32 eq.), and variation of icat/ip values in DMF solutions containing different concentrations of TFA (d).
Catalysts 14 00005 g005
Figure 6. CVs of metal corroles ((ac), 1.0 mM) in different concentrations of TsOH (0–32 eq.), and variation of icat/ip values in DMF solutions containing different concentrations of TsOH (d).
Figure 6. CVs of metal corroles ((ac), 1.0 mM) in different concentrations of TsOH (0–32 eq.), and variation of icat/ip values in DMF solutions containing different concentrations of TsOH (d).
Catalysts 14 00005 g006
Scheme 2. Suggested pathways for the Fe, Co, and Cu corrole catalysts in electrocatalytic HER.
Scheme 2. Suggested pathways for the Fe, Co, and Cu corrole catalysts in electrocatalytic HER.
Catalysts 14 00005 sch002
Figure 7. CVs of metal corroles (5.0 μM) in aqueous neutral medium.
Figure 7. CVs of metal corroles (5.0 μM) in aqueous neutral medium.
Catalysts 14 00005 g007
Table 1. Redox potentials of metal corroles.
Table 1. Redox potentials of metal corroles.
ComplexMIII/MIIMII/MI
Ox 1/VRed 1/VRed 2/V
PFIC-Co −0.43−0.84−1.97
PFIC-Cu −0.21−0.30−2.21
PFIC-Fe −0.89−1.05−2.14
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, L.-W.; Yao, Y.-F.; Xu, S.-Y.; Cao, X.-Y.; Ren, Y.-W.; Si, L.-P.; Liu, H.-Y. Electrocatalytic Hydrogen Evolution of Transition Metal (Fe, Co and Cu)–Corrole Complexes Bearing an Imidazole Group. Catalysts 2024, 14, 5. https://doi.org/10.3390/catal14010005

AMA Style

Wu L-W, Yao Y-F, Xu S-Y, Cao X-Y, Ren Y-W, Si L-P, Liu H-Y. Electrocatalytic Hydrogen Evolution of Transition Metal (Fe, Co and Cu)–Corrole Complexes Bearing an Imidazole Group. Catalysts. 2024; 14(1):5. https://doi.org/10.3390/catal14010005

Chicago/Turabian Style

Wu, Ling-Wei, Yan-Fang Yao, Shi-Yin Xu, Xu-You Cao, Yan-Wei Ren, Li-Ping Si, and Hai-Yang Liu. 2024. "Electrocatalytic Hydrogen Evolution of Transition Metal (Fe, Co and Cu)–Corrole Complexes Bearing an Imidazole Group" Catalysts 14, no. 1: 5. https://doi.org/10.3390/catal14010005

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop