Xiaojuan
Lv
a,
Long
Qian
a,
Nikolay V.
Tkachenko
*bc,
Tao
Zhang
a,
Fengxian
Qiu
*a,
Naoki
Aratani
*d,
Takahisa
Ikeue
e,
Jianming
Pan
a and
Songlin
Xue
*a
aSchool of Chemistry and Chemical Engineering, Jiangsu University, 301 Xuefu Road, Zhenjiang 212013, China. E-mail: fxqiu@ujs.edu.cn; slxue@ujs.edu.cn
bDepartment of Chemistry, University of California, Berkeley, CA 94720, USA
cMaterials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA. E-mail: nikolaytkachenko@berkeley.edu
dDivision of Materials Science, Nara Institute of Science and Technology, 8916-5 Takayama-cho, Nara 630-0192, Japan. E-mail: aratani@ms.naist.jp
eDepartment of Materials Chemistry, Graduate School of Natural Science and Technology Shimane University, 1060 Nishikawatsu, Matsue, Shimane 690-8504, Japan
First published on 27th September 2024
A novel rosarin di-Cu complex 2Cu-1 and a linear six-pyrrolic mono-copper complex 1Cu-1 were synthesized using rosarin as the ligand. The molecular conformations of these complexes were confirmed by X-ray crystallography. The optical study of 1Cu-1 indicated NIR-II absorption due to the long six-pyrrolic ligand and the ICT effect. The 2Cu-1 complex exhibited a very narrow electronic reduction–oxidation gap of 0.50 eV, attributed to the antiaromatic characteristics of the rosarin ring. The first HER study of the di-copper rosarin complex 2Cu-1 indicated that the multi-metal poly-pyrrolic complexes are promising molecular hydrogen evolution reaction catalysts.
In 2004, Osuka's group reported a series of meso-hexaphyrin(1.1.1.1.1.1) dicopper complexes (Fig. 1). The copper coordination in these complexes cleaves the global aromatic ring of meso-hexaphyrin(1.1.1.1.1.1), with O and Cl atoms serving as linked bridges between di-copper ions.9a In 1996 and 2007, Lynch and Sessler reported two di-copper complexes of hexaphyrin(1.0.0.1.0.0) and hexaphyrin(1.0.1.0.0.0), which are regarded as ring contracted hexaphyrins compared to hexaphyrin(1.1.1.1.1.1) (Fig. 1). These two di-copper complexes retained the global aromatic properties of their respective hexaphyrins, with di-copper sharing two Cl atoms as linked bridges.9b In 2019, we reported a novel figure-of-eight shaped dicopper hexaphyrin(2.1.2.1.2.1) complex. Due to the flexible structure of hexaphyrin(2.1.2.1.2.1), copper ions can control the aromaticity and cis-/trans-isomerization of hexaphyrin(2.1.2.1.2.1).5 Over the past years, we have been studying expanded porphyrins extensively.10 In 2022, we first reported highly planar hexaphyrin(1.0.1.0.1.0) (rosarin).11 We found that it can form mono-Re,12 di-Rh,4b tri-Rh,4b mono-B,13 and hybrid di-Rh/B13 complexes. These reported complexes of rosarin showed special and interesting properties, such as a highly stable antiaromatic ring,11 stabilization of the neutral radical through hetero-bimetal coordination,13 and conflicting aromaticity in trirhodium(I) rosarin.4b
In this work, we report a novel rosarin di-Cu complex (2Cu-1) and a linear six-pyrrolic copper complex (Cu-1) synthesized through a copper coordination reaction between antiaromatic rosarin (1) and copper salts. The molecular structures, optical and electronic properties, and chemical bonding of Cu-1 and 2Cu-1 were investigated using high resolution mass spectrometry (HR-MS), electron paramagnetic resonance (EPR) spectroscopy, nuclear magnetic resonance (NMR) spectroscopy, X-ray crystallography, UV-Vis absorption, electrochemistry and theoretical calculations. Furthermore, due to the presence of two Cu ions, the hydrogen evolution reaction (HER) ability of 2Cu-1 was evaluated.
Suitable single crystals of Cu-1 and 2Cu-1 were grown by slow evaporation of CH2Cl2 solution (Fig. 2 and crystal data in the ESI†). In Cu-1, the macrocyclic antiaromatic ring was cut by two –MeO groups of 1 to form a linear structure.14 Two of the three dipyrrin units in Cu-1 are coordinated with one copper ion, while the third dipyrrin unit remains uncoordinated (Fig. 2a). The copper(II) ion of Cu-1 is coordinated to four nitrogen atoms of two dipyrrin units in a distorted square planar geometry. The angle between the coordinated two dipyrrin units is 25.5°. The single crystal of 2Cu-1 exhibits a disorder involving three copper positions (Fig. 2 and crystal data in the ESI†). The free-base 1 exhibited good planarity,11 whereas the main rosarin skeleton of 2Cu-1 is slightly distorted. The mean plane deviation (MPD, defined by the 33 core C and N atoms) of 2Cu-1 reaches 0.23 Å. 2Cu-1 has a significant bond-length alternation, which is consistent with the antiaromatic nature of rosarin.11 The bond-length differences between neighbouring Cmeso–Cα bonds in 2Cu-1 are 0.050, 0.033, and 0.047 Å (Fig. 2b). Furthermore, the Cu–Cu distance in 2Cu-1 is longer than the typical Cu–Cu bond in the Cu2 dimer (2.26 Å), indicating weak or no interaction between copper ions in 2Cu-1.15
An attempt was made to record the 1H NMR spectrum for 2Cu-1. The spectrum revealed two very broad peaks corresponding to β-pyrrolic H-atoms, indicating the open-shell structure of 2Cu-1 (Fig. S4, ESI†). Thus, EPR analysis is required to reveal the nature and features of the 2Cu-1 open-shell system (Fig. S5, ESI†). Given the even number of electrons, we expect that at least two unpaired electrons are present in the structure. The EPR spectrum of 2Cu-1 exhibited g values of 2.08 and 2.29. The observed EPR spectrum is in good agreement with the simulated spectrum with the spin Hamiltonian parameters: S = 1, g⊥ = 2.08, g∥ = 2.29, D = 5.01 × 10−3 cm−1, E = 1.67 × 10−3 cm−1, A⊥(Cu) = 5.00 × 10−4 cm−1, A∥(Cu) = 8.00 × 10−3 cm−1, A⊥(N) = 3.50 × 10−3 cm−1, and A∥(N) = 2.00 × 10−3 cm−1.16 This simulated result indicates that the two unpaired electrons effectively exhibit zero coupling between each other.17
To reveal the nature of Cu–Cu interactions within 2Cu-1 and analyze its spin density distribution, we performed hybrid DFT analysis of the synthesized complex. Geometries for different spin states (singlet, triplet, and quintet) were optimized to reveal the energetically lowest spin state. Upon optimization, the geometries for different spin states were found to be mainly similar, retaining the main structural features. The calculated Cu–Cu distances for singlet, triplet, and quintet states are 2.36 Å, 2.34 Å, and 2.86 Å, respectively. The most energetically stable spin state was found to be the triplet state, which is lower in energy by 25.4 kcal mol−1 compared to the closed-shell singlet state and 20.5 kcal mol−1 compared to the open-shell quintet state. The computed Cu–Cu distance for the ground triplet state is slightly shorter than the experimentally obtained Cu–Cu distance for 2Cu-1, which can be attributed to the underestimation of the metal–metal bond distance with hybrid functionals. Further optimization of the structure using the GGA PBE method resulted in a Cu–Cu distance of 2.41 Å, which shows better agreement with the experimental values.
The unpaired spin density for the optimized triplet structure is shown in Fig. 3a. As observed, the spin density is significantly delocalized with the major density spreading to the rosarin ligand and Cu2 parts. By integrating the spin densities, it is evident that approximately one electron is localized on the Cu2 fragment, while the other electron resides on the rosarin part. Similar spin density delocalization was observed before in copper-containing compounds.17 In the geometry of triplet species, we also calculated the broken-symmetry open-shell singlet to understand the spin-coupling constant between two unpaired electrons. The calculated spin-coupling constant was found to be close to 0 eV, which agrees with the EPR results.
Because the Cu–Cu distance of 2Cu-1 is longer than a typical Cu–Cu bond,15 we decided to decipher the chemical bonding pattern of 2Cu-1 using the Adaptive Natural Density Partitioning (AdNDP) localization scheme. The AdNDP and NBO analyses revealed four lone pairs on each Cl atom, formally giving them a charge of −1. Additionally, five alpha d-type electrons and four betta d-type electrons were localized on each Cu atom. However, beta electron density partitioning reveals the presence of an additional 2c–1e Cu–Cu σ-bond with an occupation number of 0.99|e|, formed by d-orbitals (Fig. 3b). This gives the Cu2 fragment a +3 formal charge (since there are 19 valence electrons on two Cu atoms, while a neutral Cu2 would have 22 electrons) and a formal average oxidation state of +1.5 for each Cu atom. This chemical bonding analysis is consistent with the observed spin-density distribution, where the Cu2 fragment has one unpaired electron (from AdNDP we have 10 alpha and 9 beta electrons on the Cu2 fragment). The rest of the chemical bonding is in agreement with the AdNDP analysis of previously studied rosarin complexes,4b with the only difference being that instead of having six delocalized 33c–2e bonds on the rosarin ligand, we have five delocalized 33c–2e bonds and the presence of an unpaired 33c–1e bond with an occupation number of 0.98|e|. The presence of 11 instead of 12 delocalized electrons in the rosarin ring makes its antiaromatic characteristics less pronounced. The QTAIM analysis confirms the presence of weak interactions between Cu atoms, revealing a bond critical point between Cu atoms with a low electron density of 0.05 a.u. and an electron localization function (ELF) value of 0.25, indicating the presence of weak bonding interactions.
The optical properties of Cu-1 and 2Cu-1 were investigated using UV-vis-NIR absorption spectra in CH2Cl2 (Fig. S6a, ESI†). The 2Cu-1 complex exhibited a main absorption band at 484 nm. Interestingly, the linear six-pyrrolic complex Cu-1 exhibited a main absorption band at 537 nm and other bands in the NIR region, at 830 and 930 nm. These absorption spectra reflect the antiaromaticity of 2Cu-1 and the open-chain tetrapyrrolic metal complex of Cu-1.11,12,18 The TD-DFT analysis of 1Cu-1 was performed to evaluate the optical absorption. The ICT of the band at 930 nm was confirmed (Fig. S7 and S8 and Table S1, ESI†).19 The redox properties of 2Cu-1 were measured using cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in CH2Cl2 (Fig. S6b, ESI†). The amount of Cu-1 is too small to perform CV and DPV. 2Cu-1 showed three reversible oxidation and two reversible reduction E1/2 potentials at 0.84, 0.43, 0.17, −0.33, and −0.56 V (vs. SCE). The cyclic voltammetry results of 2Cu-1 show that the first oxidation (Eox1) and first reduction (Ered1) E1/2 potentials gave a very narrow gap of 0.50 eV (Egap, cv = Eox1 − Ered1).10–12 Based on our previous work,11–13 these oxidation–reduction peaks are attributed to the macrocycle itself. The DFT calculated SOMO(α)–SUMO(β) gaps of 2Cu-1 are 1.44 and 1.65 eV, respectively (Fig. S9, ESI†). The relatively narrow calculated SOMO(α)–SUMO(β) gaps are due to the antiaromatic rosarin ring.1,3,11
The 2Cu-1 was loaded on a carbon nanotube (CNT) to obtain the HER ink 2Cu-1@CNT to investigate its heterogeneous electrocatalytic HER capacity in 0.5 M H2SO4 aqueous solutions at room temperature.10c The HER ink morphology of 2Cu-1@CNT was characterised by using scanning electron microscopy (SEM) and mapping images (Fig. S10, ESI†), which showed uniform distribution of 2Cu-1 on the CNT. The typical Cu signals of 2Cu-1@CNT were confirmed by X-ray photoelectron spectroscopy (XPS) (Fig. S11, ESI†). The high-resolution XPS spectra of Cu 2p1/2 and Cu 2p3/2 showed peaks at 936.1 and 953.2 eV, respectively, corresponding to Cu ions, which is similar to the XPS result of the porphyrin(1.1.1.1) Cu complex and the benzene-fused porphyrin(2.1.2.1) complex loaded on CNTs.10c The SEM and XPS results indicated that we successfully prepared the HER ink. The linear sweep voltammograms (LSVs) of 2Cu-1@CNT show that it was more HER active in aqueous solutions (Fig. 4a). 2Cu-1@CNT gave a lower overpotential (550 mV) than the porphyrin(2.1.2.1) copper, benzene-fused porphyrin(2.1.2.1) di-copper and porphyrin(1.1.1.1) copper complexes to actuate a current density of 10 mA cm−2 under the same HER measurement conditions.10c,20 The LSV result of 2Cu-1@CNT is evidence that 2Cu-1 is a good di-copper porphyrinoid heterogeneous HER catalyst.10c,20 The ideal HER activity of 2Cu-1 is due to di-Cu ions and strong π–π interactions between the rosarin larger π-ligand of 2Cu-1 and the CNT.10c,21 The Tafel plot of 2Cu-1@CNT is 190 mV dec−1 (Fig. 4b). The Nyquist plot of 2Cu-1@CNT has a semicircular shape, and the diameter of the semicircular is smaller than that of copper porphyrinoid heterogeneous HER catalysts, indicating that 2Cu-1@CNT has a lower charge transfer resistance (Rct) (Fig. 4c). CV was conducted at different scan rates (Fig. 4d) to estimate the electrochemically active surface area (ECSA) by measuring Cdl (Fig. S12, ESI†). The Cdl value of 2Cu-1@CNT is 1.5 mF cm−2, indicating that 2Cu-1@CNT exhibits high HER catalytic activity as a molecular catalyst. The HER catalytic stability of 2Cu-1@CNT was demonstrated using an I–t test (Fig. S13, ESI†). 2Cu-1@CNT was shown to be stable during electrolysis in 0.5 M H2SO4 for the HER for 12000 s. The stability of 2Cu-1 after the HER catalysis was also investigated by UV-Vis spectroscopy. After catalysis, 2Cu-1 was obtained using ultrasound at the electrode. The absorption spectrum of 2Cu-1 after HER catalysis supports the good stability of 2Cu-1 (Fig. S14, ESI†). The first HER study attempt of the di-copper hexaphyrin complex indicated that the multi-metal ploy-pyrrolic complexes are promising molecular HER catalysts.
Fig. 4 (a) Blank LSV curves (red line) and HER LSV curves (black line) of 2Cu-1 in 0.5 M H2SO4, (b) Tafel slope plot, (c) Nyquist plot and (d) CVs at different scan rates of 2Cu-1@CNT. |
Footnote |
† Electronic supplementary information (ESI) available: Instrumentation and materials, detailed synthesis, theoretical calculations, supporting figures, and crystal data. CCDC 2365116 and 2365126. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt02161d |
This journal is © The Royal Society of Chemistry 2024 |