Open Access Article
Tao
Jiang†
a,
Yuanbo
Zhou†
b,
Tao
Ye†
c,
Shihao
Liu
b,
Ke
Li
b,
Xiaohui
Zhao
*b,
Zhongti
Sun
cd,
Naoki
Aratani
e,
Hiroko
Yamada
f,
Fengxian
Qiu
*a,
Jianming
Pan
*a,
Toshiharu
Teranishi
f and
Songlin
Xue
*a
aSchool of Chemistry and Chemical Engineering, Jiangsu University, 301 Xuefu Road, Zhenjiang 212013, China. E-mail: slxue@ujs.edu.cn
bSchool of Optical and Electronic Information, Jiangsu/Suzhou Key Laboratory of Biophotonics, International Joint Metacenter for Advanced Photonics and Electronics, Suzhou City University, Suzhou 215104, China
cSchool of Materials Science and Engineering, Jiangsu University, 301 Xuefu Road, Zhenjiang 212013, China
dAnhui Province Key Laboratory of Efficient Conversion and Solid-State Storage of Hydrogen & Electricity, Anhui University of Technology, 243002, Maanshan, China
eDivision of Materials Science, Nara Institute of Science and Technology, 8916-5 Takayama-cho, Nara 630-0192, Japan
fInstitute for Chemical Research, Kyoto University, Gokasho, Uji, Kyoto 611-0011, Japan
First published on 20th October 2025
The synthesis of wave-shaped porphyrin(2.1.2.1) arrays, including trimers and tetramers, was achieved for the first time through substitutional effects and building blocks. This enabled us to reveal the hydrogen evolution reaction (HER) capacity of the wave-shaped Cu(II) porphyrin(2.1.2.1) arrays. The trinuclear Cu(II) porphyrin(2.1.2.1) array exhibits remarkable HER capacity due to three catalytic Cu(II) ions and a π-ligand of porphyrin(2.1.2.1) array, resulting in lower overpotential, faster HER and charge transfer kinetics, and longer catalytic and charge transfer stability. In this study, we present a novel approach for the development of electrocatalytic molecular catalysts with multiple metal centres.
![]() | ||
| Fig. 1 (a) Examples of HER porphyrin mono-/multi-metal complexes; (b) our designed synthesis strategy for wave-shaped Cu(II) porphyrin arrays. | ||
In recent years, molecular catalysts containing two or more metal centres have attracted considerable interest due to their synergistic effects and multiple catalytic centres.4 In 2018, Moore et al. reported that dinuclear Cu(II) fused porphyrins (Fig. 1a) exhibited enhanced hydrogen production with near-unity faradaic efficiency and a maximum turnover frequency exceeding 2
000
000 s−1.4a Following this, in 2020 and 2023, the Apfel and Sarkar groups also reported other dinuclear Ni(II)4b and dinuclear Cu(II)4c porphyrins as HER molecular catalysts (Fig. 1a). However, to the best of our knowledge, molecular catalysts containing three or more metal catalytic centres have not yet been studied.
In our previous work, we reported the synthesis of dinuclear Cu(II) porphyrin (2.1.2.1) arrays using DPB and TPB (Scheme 1a) as building blocks.5 However, we could not find any longer Cu(II) porphyrin (2.1.2.1) arrays, so we investigated the solubility of longer Cu(II) porphyrins. In this work, we focused on using the substitutional effect and a building block strategy to synthesize wave-shaped Cu(II) porphyrin(2.1.2.1) arrays (Fig. 1b). First, we report the design and synthesis of trinuclear Cu(II) porphyrin(2.1.2.1) arrays in ideal yields through simple condensation and coordination reactions. The molecular structure of the obtained porphyrin(2.1.2.1) arrays and their Cu(II) complexes was confirmed using high-resolution mass spectrometry (HR-MS), nuclear magnetic resonance (NMR) spectroscopy, electron paramagnetic resonance (EPR) spectroscopy and X-ray crystallography. Their optical properties and electronic structures were revealed using UV-Vis absorption and density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations. Due to the presence of three catalytic Cu(II) ions, the trinuclear Cu(II) porphyrin arrays exhibit remarkable HER activity: lower overpotential, faster HER and charge transfer kinetics, and longer catalytic stability than mono- and dinuclear Cu(II) porphyrin complexes.3,5
Ligands 2P, 3P and 4P were characterized using HR-MS spectra. The HR-MS spectra of 2P, 3P and 4P exhibit the corresponding molecular ion peaks at m/z = 1322.6709 (calcd for C94H82N8 = 1322.6662 [M]+) and m/z = 1944.9749 (calcd for C138H120N12 = 1944.9758 [M]+) and m/z = 2567.2697 (calcd for C182H158N16 = 2567.2855 [M]+) (Fig. S1–S3). As 4P is present in trace amounts, the molecular structures of 2P and 3P were investigated by NMR spectroscopy. The 1H NMR spectra of 2P and 3P reflect non-global aromaticity due to their highly bent structures. Both 2P and 3P exhibit two sets of β-H resonances between 6.6 and 6.2 ppm corresponding to the pyrrole units, and inner NH protons were observed at approximately 12.2 ppm (Fig. S4 and S5). The molecular structure of 2P was determined using X-ray crystallography (Fig. 2). Unlike the reported planar and bent porphyrin(1.1.1.1) arrays,6 the crystal structure of 2P showed a wave-shaped molecular structure with two porphyrin(2.1.2.1) units. In 2P, the two porphyrin(2.1.2.1) units have a bent structure similar to that of 1P.5 Despite multiple attempts, single crystals of 3P and 4P suitable for testing could not be obtained, likely due to their excessively large molecular structures. The molecular structures of 3P and 4P were investigated using DFT calculations and were found to be wave-shaped, similar to 2P (Fig. S6 and S7).
The synthesized mono-porphyrin (1P) and wave-shaped porphyrin arrays (2P and 3P) were reacted with Cu(OAc)2·H2O in a CHCl3/MeOH mixture solution under N2 at room temperature for one hour, forming the metal complexes 1PCu, 2PCu and 3PCu with isolated yields of over 90% (Scheme 1b and the Experimental section in the SI).5 Characterization of these Cu(II) complexes revealed that the mono porphyrin 1P and the wave-shaped porphyrin arrays 2P and 3P can act as tetrapyrrolic macrocyclic ligands. The molecular structures of 1PCu, 2PCu and 3PCu were determined using HR-MS spectrometry and EPR spectroscopy. The HR-MS spectra of 1PCu, 2PCu and 3PCu exhibit the corresponding molecular ion peaks at m/z = 762.2759 [M + H]+ (calcd for C50H43CuN4 = 762.2706 [M + H]+), m/z = 1445.4961 [M + H]+ (calcd for C94H79Cu2N8 = 1445.4942 [M + H]+), and m/z = 2129.7170 [M + H]+ (calcd for C138H115Cu3N12 = 2129.2502 [M + H]+) (Fig. S8–S10). The Cu(II) ions in 1PCu, 2PCu and 3PCu were confirmed by EPR spectroscopy in a CH2Cl2 solution, which showed clear Cu(II) signals (Fig. S11–S13 and 3).3–5
![]() | ||
| Fig. 3 UV-vis absorption spectra of 1PCu (black line), 2PCu (red line) and 3PCu (blue line) in CH2Cl2 at 293 K. | ||
UV-Vis absorption spectroscopy was employed to examine the optical and electronic properties of wave-shaped porphyrin arrays and their copper complexes. The UV-Vis absorption spectra of 2P and 3P were measured in a chloroform solution. Both 2P and 3P show a similar absorption profile to 1P: they both exhibit a main absorption band at 450 and 455 nm, respectively (Fig. S14).51PCu and 2PCu both exhibited a red-shifted main absorption band at 489 and 525 nm respectively, compared to the main absorption bands of 1P and 2P. Interestingly, 3PCu exhibited two red-shifted main absorption bands at 502 and 542 nm, resembling a combined absorption spectrum of 1PCu and 2PCu.5 To gain a deeper understanding of the absorption and electronic structure of 3PCu, DFT and TD-DFT calculations were carried out. Considering the frontier molecular orbitals, the TD-DFT calculations predicted that the absorption of 3PCu would exhibit intramolecular charge transfer (ICT) characteristics.7 For 3PCu, TD-DFT predicted that the main absorption peaks at 550.41 nm (f = 0.0353) and 506.54 nm (f = 0.2846) correspond to β338 to β341 and β340 to β343 and α339 to α342, β300 to β347 and β300 to β348, respectively (Fig. S15–S17 and Table S1).
Two and three Cu centres of 2PCu and 3PCu have attracted our attention to investigate their catalytic properties.2–4 A series of experiments were carried out using a three-electrode system in 1.0 M KOH to evaluate the alkaline HER activity of 1PCu@CNT, 2PCu@CNT and 3PCu@CNT. Prior to the electrochemical performance tests, the 1PCu@CNT, 2PCu@CNT and 3PCu@CNT samples were mixed with carbon nanotubes (CNTs) and loaded onto carbon cloth.5a The morphological and compositional characteristics of the samples were systematically investigated using scanning electron microscopy (SEM) and elemental mapping analysis (Fig. 4 and S18–S20). SEM images confirmed the formation of uniformly dispersed nanostructures on CNTs. Elemental mapping analysis revealed homogeneous spatial distributions of carbon (C), nitrogen (N), and copper (Cu) across the substrates.
![]() | ||
| Fig. 4 (Left) SEM images and (right) EDS elemental mappings (Cu-L) of (a) 1PCu@CNT, (b) 2PCu@CNT and (c) 3PCu@CNT. | ||
High-resolution X-ray photoelectron spectroscopy (XPS) analysis of 1PCu@CNT, 2PCu@CNT and 3PCu@CNT revealed characteristic Cu 2p peaks at binding energies of 935.1 ± 0.2 eV (Cu 2p3/2) and 955.2 ± 0.3 eV (Cu 2p1/2), as shown in Fig. 5. The XPS spectrum definitively identified carbon (C), nitrogen (N) and copper (Cu) as the constituent elements, showing close agreement with literature reports for Cu(II)-porphyrin(1.1.1.1)/CNT hybrid systems and Cu(II)-porphyrin(2.1.2.1)/CNT data. This confirms the stable +2 oxidation state of the copper centres.5a
![]() | ||
| Fig. 5 (a) XPS survey spectra of 1PCu@CNT, 2PCu@CNT and 3PCu@CNT and (b) the corresponding Cu 2p binding energy regions. | ||
Linear sweep voltammetry (LSV) of 1PCu@CNT, 2PCu@CNT, and 3PCu@CNT was then conducted. As displayed in Fig. 6a, the HER activity of 2PCu@CNT and 3PCu@CNT catalysts all outperformed that of the 1PCu@CNT catalyst. Among them, the 3PCu@CNT catalyst exhibited the lowest overpotential to reach a current density of 10 mA cm−2, which is 137 mV lower than that of the 1PCu@CNT catalyst and 95 mV lower than that of the 2PCu@CNT catalyst. And such activity also stands out among most of the metallic porphyrin-based catalysts.5a,8,9 The HER activity is related to the strong π–π interaction between the extended conjugated skeleton of the porphyrin array, which can not only stabilize the catalyst surface, but also accelerate charge transfer.5a,8,10 So the HER activity gradually increases as the porphyrin Cu(II) unit increases from 1 to 3.
![]() | ||
| Fig. 6 (a) HER LSV curves of catalysts in 1.0 M KOH. (b) Tafel slope plots. (c) Nyquist plots. (d) i–t curves of 3PCu@CNT in 1.0 M KOH at −0.6 V vs. RHE. | ||
To further probe the HER kinetics, the Tafel curves of 1PCu@CNT, 2PCu@CNT, and 3PCu@CNT were plotted. As shown in Fig. 6b, the Tafel slopes of 2PCu@CNT and 3PCu@CNT are all smaller than that of 1PCu@CNT (301 mV dec−1). Among them, 3PCu@CNT exhibits the smallest Tafel slope (265 mV dec−1), proving its fastest HER kinetics. Electrochemical impedance spectroscopy (EIS) was also conducted to evaluate the conductivity of the circuit. As displayed in Fig. 6c, all the Nyquist plots of these samples exhibit a semicircle shape. Among them, the radius of 3PCu@CNT is the smallest, which indicates its lowest charge transfer resistance (Rct) and also proves its fastest charge transfer kinetics. In addition, cyclic voltammetry (CV) was performed at various scan rates to evaluate the double layer capacitance (Cdl). As exhibited in Fig. S21–S26, the Cdl values of 1PCu@CNT, 2PCu@CNT, and 3PCu@CNT are 0.59 mF cm−2, 0.72 mF cm−2, and 1.34 mF cm−2, respectively, indicating that 3PCu@CNT possesses the highest electrochemical active surface area. Meanwhile, 3PCu@CNT also maintained outstanding performance with no obvious current density change during 72
000 seconds of the HER stability test in 1.0 M KOH (Fig. 6d).
To further elucidate the difference in HER activity of wave-shaped Cu(II) porphyrin arrays, we use curved (C-Cu) and plane Cu (P-Cu) porphyrin complexes as catalytic units (Fig. S27) to do theoretical calculation. The first-principles method was executed to calculate the adsorption energy of hydrogen (ΔG*H),11–14 and computational details are displayed in the SI. The optimized C-Cu and P-Cu molecule models are shown in Fig. 7a and b, and the corresponding adsorbed hydrogen configurations are indicated in Fig. 7c–e. Their lowest adsorption energies of hydrogen are 1.84 and 2.03 eV on the C-Cu and P-Cu catalysts, respectively, manifesting stronger H uptake on C-Cu than on P-Cu. The C-Cu molecule catalyst also exhibits shorter bond length between H and Cu atoms with 1.62 Å than P-Cu with 1.68 Å. Fig. 7f shows the free energy change of the HER. The C-Cu molecule catalyst demonstrates lower free energy change than P-Cu by 0.19 eV, indicating preferred HER performance. What's more, they possess the same potential determining step (PDS) with the capture of protons. Above all, the C-Cu molecule catalyst is significantly beneficial for the HER. A comparative table summarizing the HER performance of the developed 3PCu@CNT catalyst alongside other recently reported porphyrin complex catalysts is given. The result indicates that 3PCu@CNT is a potential molecular HER catalyst (Table S2).
CCDC 2490044 contains the supplementary crystallographic data for this paper.15
Supplementary information is available. See DOI: https://doi.org/10.1039/d5sc07383a.
Footnote |
| † These authors contributed equally. |
| This journal is © The Royal Society of Chemistry 2025 |