Xunwei
Ma‡
ab,
Yifan
Zhang‡
a,
Liugang
Wu
a,
Zijun
Huang
a,
Jiyuan
Yang
d,
Chunguang
Chen
a,
Shengwei
Deng
e,
Lincai
Wang
*b,
Jian
Chen
*c and
Weiju
Hao
*a
aUniversity of Shanghai for Science and Technology, Shanghai 200093, P. R. China. E-mail: wjhao@usst.edu.cn
bSchool of Resources and Environmental Engineering, Shanghai Polytechnic University, Shanghai 201209, P. R. China. E-mail: lcwang@sspu.edu.cn
cDepartment of Orthopedics, Shanghai General Hospital, Shanghai Jiaotong University School of Medicine, Shanghai Jiaotong University, Shanghai 200080, China. E-mail: chenjianpumch@163.com
dDepartment of Materials Science and Engineering, National University of Singapore, Singapore
eCollege of Chemical Engineering, Zhejiang University of Technology, Hangzhou, Zhejiang 310014, P. R. China
First published on 6th November 2024
The development of highly active and cost-effective catalytic electrodes that function effectively across a wide range of pH values is one of the challenges to achieving efficient and stable hydrogen production via electrolytic water. This work constructs a self-supported catalytic electrode (Pt-NiB@NF) by growing boron-based catalytic materials in situ on nickel foam (NF) through mild electroless plating and then rapidly “decorating” trace amounts of platinum (Pt) on the precursor surface via electrodeposition. Decorating with trace amounts of Pt (0.58 wt%) achieves a 3.5-fold enhancement in the performance of NiB@NF. Pt-NiB@NF exhibits low hydrogen evolution reaction (HER) overpotentials of 70 mV and 12 mV at a current density of 100 mA cm−2 in neutral high-salt media and alkaline environments, respectively. Meanwhile, Pt-NiB@NF demonstrates long-term stability at industrial-scale current densities, maintaining for 120 hours at 100 mA cm−2 in neutral high-salt media and for 1200 hours at 500 mA cm−2 in alkaline electrolyte. The strategy of mild electroless plating and rapid electroplating realizes large-area electrode preparation for assembling a proton exchange membrane electrolyzer, more promising for industry-grade hydrogen production via water splitting. This work provides an optimized solution for the commercialization and large-scale production of high-performance Pt-based electrodes through a simple preparation strategy.
Transition metal borides (TMBs) are potential materials for HER catalytic electrodes because of their adjustable structure, high selectivity, good conductivity, excellent corrosion resistance and low cost.12 For example, Wang et al. used pulsed laser deposition-assisted synthesis of MoB–Ni3B for hydrogen production by alkaline (1.0 M KOH) and acidic (0.5 M H2SO4) water splitting to reach a current density of 10 mA cm−2 with overpotentials of 35 mV and 57 mV, respectively.13 Although boron-based materials showed promising potential in promoting water splitting, the problems of complex synthesis processes, high overpotentials and poor long-term stability faced under industrial-scale conditions need to be further addressed.14 Studies showed that surface or interfacial engineering can alter the catalytic electrode active sites and modulate the adsorption behaviour of reaction intermediates to achieve enhanced catalytic activity.15 In particular, designing metal–carrier interactions can effectively modulate the electronic structure and enhance the kinetics of the HER.16,17 Lau et al. constructed built-in electric field (BEF) metal-supported interfacial electrocatalysts Pt@CoOx, revealing that a strong BEF induced by the work function (ΔΦ) optimizes the hydrogen adsorption free energy (ΔGH*) and hydroxide adsorption free energy (ΔGOH*), thereby promoting neutral HER kinetics.18 Qiu and Cheng et al. fabricated a RuIr@BCN catalytic electrode by loading iridium (Ir) and ruthenium (Ru) nanoalloys onto boron/nitrogen co-doped graphite nanotubes (BCNs) through pyrolysis, which exhibited excellent alkaline HER activity with an overpotential of 23.6 mV at a current density of 10 mA cm−2.19 Platinum group metals (PGMs) such as platinum (Pt), iridium (Ir), and ruthenium (Ru) exhibit excellent catalytic performance.20–22 Constructing metal–carrier heterostructures of PGM-TMB opens new possibilities to enhance the catalytic activity and industrial-grade stability of the HER. However, the scarcity and high cost of PGMs have hindered their large-scale application.23 Meanwhile, achieving the real utilization of catalytic electrodes remains a challenge with the core issue being the design of composite advantage electrodes that combine “High stability, Durability, and Cost-effectiveness”. Therefore, to ensure the catalytic electrode's ultra-high activity and stability, efforts should be made to maximize the utilization efficiency of platinum group metals to reduce the content of platinum group metals in the catalytic electrode.
Furthermore, catalytic electrodes that can efficiently catalyze water splitting in a wide range of pH media are more promising for industrialization and commercialization. In particular, the efficient and stable neutral salt water splitting is beneficial for: (1) directly splitting seawater and high-salt wastewater, or (2) coupling HER withthe chlor-alkali process, which facilitates green hydrogen production in multiple scenarios and demonstrates greater commercial value.24,25 However, seawater and high-concentration salt water contain a large number of chloride ions (Cl−), and when the electrode materials are exposed to these electrolytes for a long time, the chloride ions will combine with the metal sites and lead to corrosion of the electrode materials.26 Moreover, with the increase of salt concentration in the electrolyte, the catalytic activity will be further reduced.27 Researchers devised methods such as site-specific adsorption as well as electrostatic shielding to reduce the effect of Cl− on water splitting.28 Shen et al. showed that the oxide (CeO2)−anion (B(OH)4−) dual-layer was effective in resisting the invasion of Cl− (0.5 M).29 The boron component on the surface of TMBs is easily oxidized to produce a borate anion layer when exposed to air, providing new possibilities for electrostatic repulsion of Cl− by TMB electrode materials.30 The challenge of directly splitting highly concentrated salt water with complex and corrosive compositions for hydrogen production lies in the preparation of cost-effective, highly active, and durable electrode materials through easy and controllable preparation strategies.
Based on the above considerations, this work used a mild electroless plating method to deposit a nickel–boron (Ni–B) catalytic material in situ on nickel foam (NF), and then quickly “decorated” it with trace amounts of Pt through electroplating, to construct efficient and stable Pt-NiB@NF self-supported catalytic electrodes. This strategy had the following advantages: (1) the nickel foam (NF) as the substrate provided a high surface area and good electron transfer ability for the catalytic electrode; (2) decorating Ni–B with Pt adjusted the electronic structure, providing excellent corrosion resistance and stability; (3) the mild electroless plating coupled with the electroplating process provided favorable conditions for preparing industrial-grade and practical electrodes. The Pt-NiB@NF catalyst achieved a HER current density of 10 mA cm−2 with only 12 mV overpotential in an alkaline environment (1.0 M KOH). Moreover, it could be durable for over 1200 hours at a high current density of 500 mA cm−2 without significant performance degradation. The catalyst reached a HER current density of 10 mA cm−2 with only 70 mV overpotential in a neutral high-salt environment (1.0 M NaCl). It also demonstrated stable operation for over 120 hours at a high current density of 100 mA cm−2. Furthermore, the large area Pt-NiB@NF electrode was applied in a proton exchange membrane electrolyzer cell (PEMEC), where it could achieve efficient hydrogen production (approximately 130 mL min−1) in both neutral and alkaline environments. This work provides a reference for designing efficient, stable, and economical HER catalytic electrodes and offers strategies for the commercialization and scaling up of high-performance Pt-based catalytic electrodes.
![]() | ||
Scheme 1 The optimal synthesis route of the Pt-NiB@NF electrode for the HER in both neutral and alkaline systems. |
The NiB@NF and Pt-NiB@NF electrodes were characterized using X-ray diffraction (XRD) (Fig. 1a). Three distinct diffraction peaks were observed at 44.7°, 52.1°, and 76.8° for both samples, corresponding to the (111), (200), and (220) crystal planes of face-centered cubic nickel foam (NF) metal (PDF#03-1051).26,31 A diffraction peak at 27.8° was observed, attributed to the (310) crystal plane of B2O3 (PDF#29-0236).32 Due to the low Pt content and highly dispersed Pt atoms, the Pt material might exist in an amorphous or weakly crystalline state, and hence no diffraction peaks corresponding to Pt compounds were observed in the XRD pattern.33 The SEM image of NiB@NF showed that dense polyhedral particles grew uniformly on the NF substrate (Fig. 1b). The porous structure provided abundant sites for Pt attachment. Compared to NiB@NF, the surface of Pt-NiB@NF was uniformly loaded with a large number of spherical particles, providing more active sites for the catalytic process (Fig. 1c). Furthermore, SEM elemental mapping confirmed the uniform distribution of Pt, Ni, B, and O elements on the surface of Pt-NiB@NF (Fig. 1d). Inductively coupled plasma optical emission spectroscopy (ICP-OES) confirmed that the mass fractions of Ni, B, and Pt in Pt-NiB@NF were 95.31%, 1.66%, and 0.58%, while the mass fractions of Ni and B in NiB@NF were 97.80% and 1.84%, respectively (Table S1†). To further study the crystalline state of the synthesized material, the surface catalytic layer was stripped from Pt-NiB@NF through ultrasonic treatment for transmission electron microscopy (TEM) analysis. As shown in Fig. 1e, particles with a diameter of 5–10 nm were uniformly attached to the surface of nanospheres with a diameter of 200–300 nm. High-resolution transmission electron microscopy (HRTEM) revealed that the surface of Pt–NiB contains many amorphous/weakly crystalline regions (Fig. 1f). In certain areas, lattice fringes with a spacing of 0.202 nm, corresponding to the (111) plane of metallic Ni, were observed (Fig. 1f1 and 2). Meanwhile, lattice fringes with a spacing of 0.226 nm were also observed dispersed at the edges of these amorphous/weakly crystalline regions, which can be attributed to the (111) plane of Pt (Fig. 1f3 and 4). It indicated that the surface-deposited Pt species are loaded onto the NiB substrate in the form of clusters or nanoparticles. Similar results were obtained after multiple repeated experiments (Fig. S9†). Additionally, the selected area electron diffraction (SAED) pattern further confirmed the crystal structure of Pt-NiB@NF, with diffraction rings matching the (111) plane of Pt and (200) and (220) planes of Ni (Fig. S10†), consistent with the results of XRD. These results indicate that the NiB substrate mainly existed in an amorphous/weakly crystalline form. This was because electroless deposition was a self-growth process, and the low temperature prevented the formation of crystalline nuclei, leading to a disordered amorphous/weakly crystalline structure, which was also reported in the literature.29,34 The electrodeposited Pt species were anchored on these amorphous/weakly crystalline NiB substrates in the form of clusters or nanoparticles.
To further analyze the surface valence states of NiB@NF and Pt-NiB@NF electrodes, X-ray photoelectron spectroscopy (XPS) was conducted. The XPS survey spectrum of Pt-NiB@NF confirmed the presence of Ni, B, O, and Pt elements (Fig. S11†). The Pt 4f spectrum of Pt-NiB@NF exhibited two main peaks at 71.9 eV and 75.1 eV, corresponding to Pt 4f7/2 and Pt 4f5/2, respectively (Fig. S12a†).35 A peak appearing at 68.49 eV overlapped partially with the binding energy peak of Pt 4f, and this peak corresponded to Ni 3p.36 In the Ni 2p spectrum of Pt-NiB@NF, the peaks located at 853.2, 855.8/873.3, 857.8/876.1, and 861.9/879.8 eV can be attributed to Ni–B bonds (Niδ), divalent Ni species (Ni2+), trivalent Ni species (Ni3+), and satellite peaks (Fig. S12b†).37,38 Compared to the Ni 2p spectrum of NiB@NF, the peaks of oxidized Ni species undergo a positive shift, while the peak of Niδ shifted positively by 0.1 eV. It indicated that the introduction of Pt results in a reduction of electron density around the Ni atoms. The peak at 192.2 eV in the B 1s spectrum was attributed to the B–O bonding, indicating that the surface boron on the electrode had been oxidized due to exposure to air (Fig. S12c†).39,40 Further observation of the O 1s spectrum of Pt–NiB shows that the peaks of surface-adsorbed water (OH2O), hydroxide (OOH), and lattice oxygen (OLattice) are located at 532.7, 531.6, and 530.9 eV, respectively (Fig. S12d†).41 Compared to the peak ratio in the O 1s spectrum of NiB@NF (OLattice/OOH = 0.6/1), that in the figure of Pt-NiB@NF increased to 0.74/1, further indicating that the introduction of Pt leads to the formation of more high-valence oxides on the surface of the catalytic material. These characterization results demonstrated the successful deposition of Pt on NiB@NF and the modulation of the intrinsic electronic structure.
The surface wettability of the NiB@NF and Pt-NiB@NF electrodes was studied by analysing the dynamic contact angle hysteresis through liquid droplet contact imaging. Fig. 1h and i show that both NiB@NF and Pt-NiB@NF exhibited good hydrophilicity. Specifically, Pt-NiB@NF achieved complete droplet penetration in just 118 ms, which was faster than NiB@NF (237 ms). Due to the large number of gas bubbles generated during the HER process at high current densities, the catalyst surface would be covered, increasing the charge and proton transfer resistance at the interface, which inhibited HER activity. Catalysts with superhydrophilicity contributed to the kinetics of the HER.42 On the other hand, an uneven, porous surface utilized the capillary effect to transport liquid electrolyte to the surface, enhancing hydrophilicity and reducing the adhesion of gas bubbles.43
The electrochemically active surface area (ECSA) was an important indicator for evaluating the catalytic performance of electrodes. In this work, cyclic voltammetry was conducted in the non-faradaic region (0.09–0.15 V vs. RHE) at different scans to obtain the double-layer capacitance (Cdl) for evaluating the ECSA (Fig. S13†). The Cdl value of the Pt-NiB@NF electrode (756.0 mF cm−2) was significantly higher than that of the NiB@NF electrode (708.6 mF cm−2) (Fig. 2d), indicating that the trace loading of Pt played a positive role in enhancing the catalytic performance of the electrode. To eliminate the influence of the electrochemically active surface area on catalytic activity, the polarization curves were normalized by Cdl. As shown in Fig. 2e, Pt-NiB@NF exhibited a higher j/Cdl value compared to NiB@NF at the same overpotential, indicating that Pt-NiB@NF possessed higher catalytic activity. To eliminate the impact of the loading amount on catalytic activity, the normalization of polarization curves by the loading amount was conducted (Fig. 2f), confirming the high HER catalytic activity of Pt-NiB@NF. By calculating the turnover frequency (TOF) of each surface site of the catalyst for H2 conversion, the intrinsic activity of all catalysts was explored. As shown in Fig. 2g, the TOF of Pt-NiB@NF was 3.46 s−1, which was 2.4 times higher than that of NiB@NF (1.45 s−1) at an overpotential of 350 mV, indicating that Pt-NiB@NF had higher intrinsic activity. The results demonstrated that the doping of Pt provided more active sites for NiB@NF, thereby enhancing the HER catalytic activity.
The long-term stability under corrosive conditions was an important indicator for industrial applications. The hydrolysis system with Pt-NiB@NF as both the cathode and the anode was continuously stabilized in a neutral high-salt medium (1.0 M NaCl) at a high current density of 100 mA cm−2 for more than 120 h with only a 9.17% increase in potential (Fig. 2h). The LSV curve for Pt-NiB@NF showed minimal changes after 120 hours of catalytic electrolysis (Fig. 2h, inset). The high catalytic activity exhibited in a neutral high-salt medium and the long-term stability at high current densities hold promise for industrial applications of Pt-NiB@NF in areas such as seawater electrolysis, high-salinity lake water electrolysis, and co-electrolysis of high-salt wastewater for hydrogen production.
The catalytic electrode was further tested for catalysis in 1.0 M KOH to expand its applicability. Fig. 3a shows the LSV curve of the electrode in 1.0 M KOH. Pt-NiB@NF still exhibited the best HER performance (η10 = 12 mV), far surpassing the commercial Pt foil (η10 = 57 mV), NiB@NF (η10 = 24 mV), and bare NF (η10 = 237 mV), respectively. Compared to recently reported HER catalytic materials, Pt-NiB@NF demonstrated significant advantages under alkaline conditions (Table S3†). The Tafel slope of Pt-NiB@NF was 48.6 mV dec−1, indicating that it corresponded to the “Volmer–Heyrovsky” catalytic process in alkaline HER.44 The Tafel slope of Pt-NiB@NF was significantly lower than that of the commercial Pt foil electrode (73.7 mV dec−1), NiB@NF (90.7 mV dec−1), and bare NF (136.4 mV dec−1) (Fig. 3b). To further confirm the rapid HER kinetics of Pt-NiB@NF, the charge transfer rate of the electrode was explored through Nyquist plots (Fig. 3c). The charge transfer resistance (Rct) of Pt-NiB@NF was 0.27 Ω, significantly lower than that of Pt foil (6.43 Ω) and NiB@NF (0.51 Ω), indicating that Pt-NiB@NF had a faster charge transfer rate.
The Cdl value of Pt-NiB@NF was 2577.6 mF cm−2, superior to that of NiB@NF (1575.6 mF cm−2) (Fig. 3d), indicating that the loading of Pt enhanced the electrochemically active surface area (ECSA) of the electrode (Fig. S14†). The j/Cdl value of Pt-NiB@NF consistently remained higher than that of NiB@NF (Fig. 3e), suggesting that after excluding the influence of the electrochemically active surface area on catalytic activity, Pt-NiB@NF still exhibited higher activity. Fig. 3f reflects the catalytic performance enhancement per unit loading amount after decorating NiB@NF with Pt, demonstrating that Pt-NiB@NF exhibited higher intrinsic catalytic activity towards the HER. The TOF value of Pt-NiB@NF consistently remained higher than that of NiB@NF (Fig. 3g), showing that the doping of Pt significantly improved the electrode's intrinsic activity. More importantly, Pt-NiB@NF exhibited extremely high stability, enduring electrolysis for over 1200 hours at an industrial-level current density of 500 mA cm−2, with only an 8.16% increase in the potential (Fig. 3h), showing virtually unchanged performance (Fig. 3h, inset).
The FT-IR spectrum elucidated the chemical structure of Pt-NiB@NF after HER stability testing (Fig. S18†). There were two peaks at 1632/3451 cm−1 corresponding to the bending and stretching vibrations of adsorbed H2O molecules, formed by absorbed water or metal hydroxides.45 The peak at 1092 cm−1 was attributed to the stretching vibration of B–O bonds. Compared to the original Pt-NiB@NF, the M–O stretching vibration at 672 cm−1 was not prominently observed in the neutral-post and alkaline-post samples. Instead, a stretching vibration peak belonging to Pt–O–Ni appeared at 745 cm−1.46 This indicated that during the HER catalytic reaction, the catalytic material might have undergone surface restructuring, resulting in the formation of a hydroxide coverage layer.
The surface composition and chemical states of Pt-NiB@NF after HER stability testing were examined by XPS. As shown in Fig. S19,† in the XPS spectra of neutral-post and alkaline-post samples, signals for Pt, Ni, B, and O elements were observed, consistent with the initial sample. In the Pt 4f spectra of the neutral/alkaline-post samples (Fig. 4a), peaks for Pt 4f7/2 and Pt 4f5/2 were observed at approximately 71.3/71.3 eV and 74.9/74.6 eV, respectively. Compared to the initial sample, the peaks of Pt 4f7/2 and Pt 4f5/2 in the neutral/alkaline-post figures shifted to lower binding energies. It may be attributed to the surface reconfiguration of the NiB support during the HER process, which interacts with the Pt metal layer resulting in electron transfer.47 Further observation of the Ni 2p spectrum in neutral-post sample reveals that Ni3+ 2p3/2, Ni2+ 2p3/2, Ni3+ 2p1/2, and Ni2+ 2p1/2 shift towards lower binding energies by 1.1, 0.5, 0.5, and 0.4 eV, respectively (Fig. 4b). Meanwhile, the peak area ratio of Ni3+/Ni2+ changes from 0.9/1.5 to 0.8/1.4. In the Ni 2p spectrum of the alkaline-HER sample, the peaks of Ni3+ 2p3/2, Ni2+ 2p3/2, Ni3+ 2p1/2, and Ni2+ 2p1/2 are located at 874.7, 872.7, 857.2, and 855.3 eV, respectively. Compared to the initial sample, the peaks of the high-valence Ni species also show a negative shift, and the peak area ratio of Ni3+/Ni2+ decreases to 0.7/1.4. This is due to the reduction of high-valence nickel species on the surface of Pt-NiB@NF to a lower valence state, resulting in the formation of nickel-based (hydroxy) oxides.48 This suggested the possible surface reconstruction of Ni(OH)2 on the samples, which might be the true active species in the HER process.49 In the B 1s spectra (Fig. 4c), the peak intensities of B–O at 192.17 eV and 192.13 eV in the neutral/alkaline-post samples, respectively, showed a sharp decrease after the HER.50 This may be due to the leaching of borate anionic elements from the surface into the electrolyte.29 The significant decrease in the intensity of the B–O peak indicated that the boron-modified NF with abundant defects could confine Ni in the Ni2+ states. This provided ample water adsorption/dissociation sites and contributed to modulating the electronic structures of Pt and Ni.51 The O 1s spectra revealed characteristic peaks of surface-adsorbed water (OH2O), hydroxide (OOH), and lattice oxygen (OLattice) at 532.3/532.1 eV, 531.0/530.8 eV, and 528.7/529.3 eV in the neutral/alkaline-post samples (Fig. 4d). Compared to the initial sample, the peak area ratios (OLattice/OOH) of hydroxide (OOH) and lattice oxygen (OLattice) for neutral-post and alkaline-post samples decreased from 0.74/1 to 0.13/1 and 0.3/1, respectively. The significant decrease in the OLattice peak ratio and the negative shift in binding energy may be related to the partial substitution of anions in metal oxides,52 possibly due to the generation of Ni(OH)2 from the reduction of high-valent nickel species during the HER process. Additionally, the negative shift of the OOH peaks may be due to changes in surface OH− adsorption sites after the catalyst material undergoes surface reconstruction, leading to a decrease in binding energy.
The in situ Raman spectroscopic characterization elucidated the surface reconstruction components of Pt-NiB@NF during the HER (Fig. 4e). The initial sample exhibited peaks at approximately 545 cm−1 and 816 cm−1, attributed to oxidized Ni and B species, respectively.53 Previous studies have shown that the anions on the electrode surface can electrostatically repel Cl− and prevent corrosion of the electrode material.26,28,29,54 Therefore, the negatively charged surface of these surface borate anions with better corrosion resistance in high-salt (1 M NaCl) electrolytes mainly originates from this electrostatic shielding effect. Additionally, upon applying a potential of −0.10 V vs. RHE in 1.0 M NaCl electrolyte, a novel, broadened Raman peak emerged near 470 cm−1, whose intensity increased with increasingly negative potentials, indicative of Ni(OH)2 species.55 Furthermore, upon imposition of a −1.0 V reduction potential, a band at 1627 cm−1 was discerned, corresponding to the H–O–H bending mode (δH–O–H) of interfacial water.56,57 These findings underscore the formation of Ni(OH)2 species on the catalyst surface during the HER process. The surface reconstruction of the NiB support leads to changes in the electronic structure. As the support for Pt, NiB acted as the “environment” for the surface Pt metal.58 Therefore, the metal–support interaction further induced electron transfer between the surface Pt metal and the support. This redistribution of charges influences the adsorption/desorption of intermediate species such as H* or OH*, thereby enhancing the HER kinetics. The specific catalytic mechanism of Pt-NiB@NF in neutral and alkaline solutions during the HER process is illustrated as Fig. 4f. During the HER, water molecules acquired an electron on the electrode surface and dissociated into OH− ions and Hads (hydrogen adsorption). The Hads reacted with water molecules and electrons to form OH− ions and hydrogen molecules (H2). The H2 accumulated on the cathode surface until it reached saturation, at which point it desorbed from the electrode surface and was released into the gas phase. Pt/Ni(OH)2/NiB served as the true active species during the HER process.
Based on the analysis above, the exceptional HER of the Pt-NiB@NF catalyst was primarily attributed to the following reasons: (i) using a three-dimensional reticular structure of nickel foam as the substrate, enabling the in situ growth of uniform and dense spherical structures, effectively providing a larger specific surface area and abundant active sites; (ii) the synergistic effect between Ni and Pt effectively modulated the material's electronic structure, optimized the adsorption energy of intermediates, exposed more active sites, and possessed higher intrinsic catalytic activity; (iii) the surface reconstruction-induced Pt/Ni(OH)2 not only synergistically promoted catalytic activity with Pt–NiB, prevented further oxidation, and enhanced electrode stability; (iv) the boron-modified NF with abundant defects was utilized to disrupt the H–OH bonds, facilitating the HER by promoting the catalytic activity of Ni and Pt.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta05770h |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2024 |