Vipin
Adavan Kiliyankil
*a,
Wei
Mao
*ab,
Yurie
Takahashi
a,
Wei
Gong
ac,
Shigeru
Kabayama
d,
Yuki
Hamasaki
e,
Katsuyuki
Fukutani
bf,
Hiroyuki
Matsuzaki
ag,
Ichiro
Sakata
ah,
Kenji
Takeuchi
c,
Morinobu
Endo
c and
Bunshi
Fugetsu
*ch
aSchool, of Engineering, The University of Tokyo, Bunkyo-Ku, Yayoi 2-11-16, Tokyo 113-0032, Japan. E-mail: vipin@ipr-ctr.t.u-tokyo.ac.jp; maowei@iis.u-tokyo.ac.jp
bInstitute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan. E-mail: fukutani@iis.u-tokyo.ac.jp
cFaculty of Engineering, Shinshu University, 4-17-1 Wakasato, Nagano-shi 380-8553, Japan. E-mail: fugetsu@shinshu-u.ac.jp; bunshifugetsu@ifi.u-tokyo.ac.jp
dNihon Trim Co. Ltd, Oyodonaka, Kita-ku, Osaka, Japan
eKyushu Electric Power Co. Inc., Technical Solution Headquarters Research Institute Low Carbon Technology Group, Minami-ku, Shiobara 2-1-47, Fukuoka City 815-8520, Japan
fAdvanced Science Research Center, Japan Atomic Energy Agency (JAEA), Naka, Ibaraki 319-1195, Japan
gThe Micro Analysis Laboratory, Tandem Accelerator, The University Museum, The University of Tokyo, 2-11-16 Yayoi, Bunkyo-ku, Tokyo 113-0032, Japan
hInstitute for Future Initiatives, The University of Tokyo, Bunkyo-Ku, Yayoi 2-11-16, Tokyo 113-8656, Japan
First published on 14th October 2024
The hydrogen evolution reaction (HER) on platinum (Pt) electrocatalysts involves the generation of hydrogen atoms and the formation of hydrogen molecules. It is commonly believed that the sites on the surfaces of the terrace (111, 110, and 100) domains are responsible for the formation of hydrogen molecules. However, the electrochemistry of the hydrogen atom generation is not well understood. We created edge-rich platinum electrocatalysts using nano-fabrics comprising entire single-walled carbon nanotubes (SWCNTs) as templates and supports. We then conducted the HER on the edge-rich Pt/SWCNT hybridized electrocatalysts and gained new insights into the electrochemical properties and functions of the edge sites. We propose that the edge sites are oxidized and serve two important functions: they act as atomic barriers, allowing electrons to accumulate within the terrace (111, 110, and 100) domains, and they transfer the electrons to the hydronium ions in the electrical double layer through discharge. Enhancing the discharge capability of the electrocatalysts is an efficient way to reduce the amount of platinum required, and this can be applied to various precious metal-based electrocatalysts to enhance their electrocatalytic activities and durability.
Pt is scarce on Earth. To reduce the amount of Pt used, it is broken down into very fine particles (Pt nanoparticles) and then loaded onto the surface of a nonmetal-based support.4 When hydrogen molecules reach or form on Pt nanoparticles, they partly dissociate into hydrogen atoms, which then move from the Pt nanoparticles to their support, a phenomenon known as hydrogen spillover. This was first observed by Khoobiar.5 Hydrogen spillover can occur quickly or slowly, depending on the physiochemical properties of the catalyst support.6 Li and colleagues investigated hydrogen spillover in an actual electrolytic HER and found that the difference in work functions between the metal-based catalyst and the nonmetal-based support caused interfacial charge accumulation, which was the main barrier to hydrogen spillover.7
Traditionally, Pt nanoparticles are loaded in isolation onto carbon granules,8,9 leaving a large uncovered area on the support's surface. Apart from delaying hydrogen spillover, the exposed surface of the support also leads to electrochemical Ostwald ripening and corrosion of the carbon-based anodic catalyst support, particularly under high current densities.10 It's worth mentioning that previous studies demonstrated the advantageous properties of carbon nanotubes in either the nitrogen-dopped11,12 or the sheet form13 as catalyst supports to strengthen the interactions between Pt and the carbon-based supports for HER, still, the Pt nanoparticles were deposited in isolation.
To address the functional barriers between catalysts and supports, van der Vliet and co-workers14 have developed support-free electrocatalysts. They used vertically aligned perylene red whiskers as templates, each coated with a nano-thin Pt or Pt alloy layer via planar magnetron sputter deposition. Afterward, the perylene red whiskers were evaporated at 400 °C in a hydrogen-rich atmosphere to create a support-free, mesostructured thin-film Pt catalyst for HER. Similarly, Higashi and colleagues established a textile IrO2-based electrocatalyst for oxygen evolution reactions (OERs) using electrospun polyvinylpyrrolidone (PVP)-based nonwoven fabrics as templates.15 The PVP-based templates were dissolved in water to produce a nonwoven fabric-shaped IrO2. However, transferring these very brittle nonwoven fabric-shaped catalysts to an ion-exchange membrane for building water electrolysis cells remains challenging.
We introduce a new type of platinum/carbon (Pt/C) hybridized electrocatalysts using nano-fabrics consisting of entire single-walled carbon nanotubes (SWCNTs) as the templates and supports. The SWCNT-based nano-fabrics have a uniform surface with specific properties, allowing for a uniformly deposited Pt layer with a thickness ranging from 1 to 20 nm on the surface of each SWCNTs. Moreover, the proximity of the SWCNTs in the nano-fabrics restricts the width of the Pt terrace domains to a few nanometres, thereby creating numerous sites on edges and steps. The HER experimental evidence shows that narrowing the Pt terrace domains eases the formation of hydrogen molecules, while increasing the ultimate number of Pt-sites on edges and steps intensifies the discharge, promoting the generation of hydrogen atoms.
Fig. 1A displays a typical scanning electron microscope (SEM) image of a highly purified SWCNT-based nonwoven nano-fabric sample. The nano-fabric sample consists of individual SWCNTs and thin bundles composed of two or a few tubes, forming a highly porous nanostructured nonwoven fabric. The density, thickness, electrical resistivity, surface area, pore diameter, micropore volume, microporous surface area, and external surface area are summarized in Table S1 of the ESI†. Additionally, Fig. 1B shows an SEM image of a highly purified SWCNT-based nano-fabric sample prepared without TEMPO-CNFs, resulting in thicker bundles of SWCNTs. The mechanical properties are summarized in Fig. S1A and Table S2 of the ESI,† while nitrogen adsorption/desorption isotherms and pore size distributions are shown in Fig. S1B and C,† respectively. The free-standing SWCNT nano-fabrics are also demonstrated to exhibit excellent flexibility and foldability, as depicted in Fig. S2 of the ESI.† Typical Raman spectra are illustrated in Fig. 1C, with the G band peak splitting into G+ at 1593 cm−1 and G− at 1572 cm−1. The corresponding band near 1339 cm−1 is the D band. The intensity ratio of the G band to the D band of the SWCNT nano-fabrics is 49, similar to the G/D ratio found for the as-received powdered SWCNTs. However, a significant change in the frequency region of 100–300 cm−1, corresponding to the radial breathing mode (RBM),24,25 is observed in Fig. 1D. The peaks associated with tubes of specific diameters were altered after carbonization in N2 and oxidation via CO2.
The X-ray photoelectron spectroscopy (XPS) was used to evaluate the elemental compositions and chemical states of the free-standing SWCNT nano-fabrics. The detailed elemental compositions are provided in Fig. S3 and Table S3 of the ESI.† The C 1s spectra were deconvoluted to determine the nature of the carbon bonding.26,27 The peaks at 284.4 and 285.6 eV correspond to sp2 (C–C bond) and sp3 (C–O bond), respectively. The percentage of sp2-hybridized carbon atoms in the refined free-standing SWCNT nano-fabrics was found to be 71.5%, which was 1.5% lower than that in the as-prepared SWCNT powders. Additionally, approximately 0.6 wt% of Fe remained in the free-standing SWCNT nano-fabrics, likely due to the residual catalyst inside the SWCNTs.
A significant decrease in the intensity of the Raman peaks of SWCNTs in the radial breathing mode (RBM) range was observed. The radial breathing of all tubes with diameters of 1.8 nm, 1.6 nm, 1.47 nm, and 1.32 nm was almost completely suppressed after the Pt layers, with thicknesses greater than 10 nm, were applied. Again, this suggests that the surface of the SWCNT-based nano-fabrics was almost entirely covered by the Pt thin layer. Fig. 3A illustrates the typical Raman intensity data in the RBM range of a free-standing SWCNT nano-fabric sample, as well as the Raman intensity data of the same SWCNT nano-fabric sample after Pt layers with the thicknesses of 4 nm, 10 nm, and 20 nm were applied.
Fig. 3B displays a high-resolution XPS Pt4f spectrum of Pt of the Pt/SWCNT hybridized nano-fabric catalysts. Two prominent peaks are observed at 71.4 eV and 74.8 eV, corresponding to Pt 4f7/2 and 4f5/2, respectively. Deconvolution of the major peaks shows two dominant metallic Pt(0) peaks at 71.4 eV and 74.8 eV, two oxidized Pt(II) peaks at 72.8 eV and 76.1 eV, and two oxidized Pt(IV) peaks at 74.2 eV and 77.4 eV.28
ESI Fig. S4† exhibits the cross-sectional SEM images and EDS mappings of Pt and carbon (SWCNTs); once again, the Pt layers are observed to be loaded in a continuous and borderless manner on the SWCNT nano-fabrics with high uniformity. The depth of Pt which can reach in the SWCNT nano-fabrics is in the range of 50–80 nm. Additionally, after loading the thin Pt layers, the surface areas of the resultant Pt/WCNT hybridized nano-fabrics were found to be as high as the surface areas of the free-standing SWCNT nano-fabric template/support. The surface areas were 1100 m2 g−1, 1050 m2 g−1, and 987 m2 g−1 for the Pt/SWCNT hybridized nano-fabrics with Pt layer thicknesses of 4 nm, 10 nm, and 20 nm, respectively. For comparison, commercially available, carbon fiber-based (carbon fiber average diameter: 8 ± 2 μm) fabrics were used as a catalyst support/template, and Pt was loaded via PVD under identical conditions. ESI Fig. S5† displays the typical SEM images. The Pt layers were found to be suspended over the carbon fiber-based fabrics and cracked into many pieces. The nearly 10 μm-sized diameters and the nonuniform (sp2/sp3) surface of the carbon fibers disrupted the formation of a continuously interconnected Pt layer on their surface.
Fig. 4 presents typical LSV curves at full scale, Tafel plots, turnover frequencies (TOFs), mass activities, and stabilities after 10000 cycles for the four Pt/SWCNT nano-fabric electrocatalysts. The data on Pt layer thickness, Pt loading ratios (μg cm−2), overpotentials at applied current densities of 10 mA cm−2, 20 mA cm−2, 50 mA cm−2, and 70 mA cm−2, TOF at an overpotential of 50 mV, Tafel slopes, and mass activities at an overpotential of 50 mV are summarized in Table S4 in the ESI.† The finest Pt/SWCNT electrocatalyst, with a 1 nm thin Pt layer (Pt loading: 1.16 μg cm−2), exhibited superior HER performance compared to the powdered Pt/C catalyst (Pt loading: 57 μg cm−2). Specifically, as the applied current density increased from 10 mA cm−2 to 70 mA cm−2, the overpotential for the Pt/SWCNT nano-fabric electrocatalyst increased by only 33 mV (from 16 mV to 49 mV), while it increased by 146 mV (from 26 to 172 mV) for the powdered Pt/C electrocatalyst. Additionally, both the TOF and mass activity at an overpotential of 50 mV were approximately two orders of magnitude greater than those of the powdered Pt/C electrocatalyst. Furthermore, the electrocatalytic activity remained almost unchanged even after 10
000 cycles. Notably, the TOF and mass activity were very sensitive to the Pt loading. As the Pt loading increased from 1.16 μg cm−2 to 61.03 μg cm−2, the TOF decreased from 64.49 to 1.59, and the mass activity decreased from 64.17 to 1.58. The Tafel slopes for the powdered Pt/C and the Pt/SWCNT nano-fabric electrocatalysts were almost identical, indicating that the resistivities of the electrons transferring from the current-collector to the cathodic catalyst at lower applied current densities were less important for the HER. Furthermore, a comparison of our Pt/SWCNT nano-fabric electrocatalyst performance with that of previously reported Pt-based electrocatalysts is provided in Table S5 in the ESI,† based on the benchmarking protocol established by McCrory and coworkers.29
Fig. 5 shows typical cyclic voltammetric profiles obtained using Pt/SWCNT hybridized nano-fabrics as the working electrode in an aqueous 0.5 M H2SO4 solution at room temperature. The upper potential limit in the scans is set at 1.40 V (vs. RHE), slightly higher than the potential for the electrochemical splitting of water. A carbon-based porous cylinder was used as the counter electrode (ESI Fig. S7†). It's important to note that a Pt-based electrode can also be used as the counter electrode, but to prevent corrosion and thereby the transfer of Pt from the counter electrode to the working electrode, the upper potential limit should always be set below 0.9 V when a Pt electrode is used as the counter electrode. The range between the terminal of the H–Pt oxidation/reduction peak (∼0.12 V) and the beginning of the oxygen-oxidation/reduction peak (∼1.40 V) corresponds to the so-called constant electrical double-layer (EDL) charging current range.30,31 The Pt/SWCNT hybridized nano-fabrics can maintain an EDL more than 3.5 times greater than that of the powdered Pt/C electrode at an identical Pt loading, as depicted in Fig. 5A, based on the electrochemical surface area (ECSA) calculation. The entire SWCNT-based nano-fabrics also showed excellent capability for maintaining the EDL and behaving entirely as an EDL capacitor. Fig. 5B shows the typical CV profiles for the first cycle and the 50th cycle. The profiles shown in Fig. 5A and B were obtained using Pt/SWCNT hybridized nano-fabrics as the working electrodes and a carbon-based porous cylinder as the counter electrode. The scan rate was optimized at 10 mV s−1. After 30–50 CV cycles, the area of the peaks due to the oxidation of Pt–H, which reflects the total number of Pt sites on the terrace (111, 110, and 100) domains capable of forming Pt–H, increased by 2.5 times and then stabilized at that level. This indicates that the Pt atoms on the surface of the Pt/SWCNT hybridized nano-fabrics self-rearranged to the most stable form during CV cycling.
To investigate the redox interactions of H–Pt, we stacked two pieces of the 20 nm-thick Pt/SWCNT nano-fabrics and used them as the working electrode to obtain the CV profiles. The CV profiles (Fig. 5C) revealed three Pt–H oxidation peaks at 0.16 V, 0.24 V, and 0.32 V, as well as three Pt–H reduction peaks at 0.31 V, 0.23 V, and 0.15 V. However, these peaks were weaker and broader compared to conventional Pt electrodes. The entire SWCNT nano-fabric also exhibited areas of the EDL in the CV profiles.
The appearance of three pairs of H–Pt redox peaks between 0.16 V and 0.32 V suggested chemical or semi-chemical bonding between hydrogen and platinum.32–34 The concentration of protons originating from Pt–H in the Pt/SWCNT hybridized nano-fabrics was estimated to be 2.5–2.8 × 1021 H-atom/Pt-cm3, indicating the presence of a certain amount of Pt–H even after the HER. This was determined using resonance nuclear reaction analysis (RNRA), a unique method for detecting protons on surfaces and within substrates.35,36 Additional details can be found in Fig. S8 of the ESI.†
Charge transfers in the HER are predominantly achieved by the catalyst support. The charge-transfer behaviours of the Pt/SWCNT hybridized nano-fabrics were evaluated by using electrochemical impedance spectroscopy (EIS). Fig. 5D shows the typical Nyquist plots obtained within the range of high frequency (105 Hz) and low frequency (0.01 Hz). Regardless of the thickness of the Pt layers on the free-standing SWCNT nano-fabrics, nearly identical shapes were observed in the Nyquist plots. The intercepts at the real axis in the high-frequency region, which reflect the sum of the ohmic resistances of the intrinsic resistance of the electrolyte, catalyst, catalyst support, contact resistance at the catalyst/support/current collector, and charge transfer resistance caused by the redox reactions, were 8.1 ohms, 8.6 ohms, and 9.2 ohms, respectively, for the Pt/SWCNT hybridized nano-fabrics with Pt thicknesses of 4 nm, 10 nm, and 20 nm, indicating the transfer of electrons from the anode to the cathodic catalyst was predominantly accomplished via the SWCNT nano-fabrics, i.e., the catalyst support. The average roughness of the surface of the SWCNT nano-fabrics was calculated to be approximately 30 nm based on the AFM image data (ESI, Fig. S9†). This finding demonstrated the advantageous properties of the SWCNT nano-fabrics as supports/templates, which minimized the contact resistance of the catalyst/support/current collector. The lines in the low-frequency region were almost vertical, indicating that the diffusion resistance of proton ions from the bulk electrolyte to the EDL was very small and could be ignored.37
The binding energies of Pt-SWCNT (14,0) and H–Pt/SWCNT (14,0) were calculated using density functional theory (DFT). Strong interactions (−2.22 eV) between Pt and SWCNT (14, 0) were observed, as well as between H and Pt/SWCNT (14,0) (−0.25 eV). Additionally, the Pt(111)/SWCNT (14,0) nano-fabrics exhibited a near-zero (−0.01 eV) Gibbs free energy, demonstrating a similar behavior to bulk Pt for hydrogen adsorption/desorption (ESI,† DFT calculations).
M–H−e + H+ ⇔ H2 + M, |
In addition to Heyrovsky's reaction, Tafel noted the following Tafel reaction; here, the combination of the hydrogen atoms to form hydrogen molecules is achieved as follows:2
M–H + M′–H ⇔ H2 + M + M′ |
It explains that two active sites, labeled M and M′, positioned nearby, on the cathodic catalyst, each has bound a hydrogen atom that is generated through the electrochemical reactions. These two hydrogen atoms then combine to form a hydrogen molecule. The Tafel reaction also means that the formation of hydrogen molecules becomes easier by narrowing the widths of the terrace domains of the electrocatalysts.
In addition, it emphasizes that regardless of Heyrovsky's reaction and/or Tafel's reaction, the crucial step in forming hydrogen molecules is the generation of hydrogen atoms on cathodic catalysts, known as the Volmer step:
{H+} + e− + M ⇔ M–H |
The H+ ions in both the electrical double layer (EDL) and in the bulk electrolyte are always combined with water, forming hydronium ions, H3O+. Therefore, the Volmer step can be represented by the following reaction, as suggested by Conway and Salomon:39
{H3O+} + 2 M + e−⇔ M·H + H2O·M |
The presence of additional hydrogen bonds (HBs) causes H3O+ to bond with more water molecules and stabilize as (H2O)4H+.40 The strength of the HBs is anisotropic, with the strongest bond calculated to be 13.8 kJ mol−1,40 approximately 5 times stronger than the HBs in bulk liquid water. The remarkably strong HBs in the hydronium ion significantly contribute to the overpotential in the HER. This challenge is addressed in our study by enhancing the discharge capability using SWCNT nano-fabrics as Pt supports.
Upon receiving an electron, the hydrated proton ion is transformed into a hydrogen atom, leaving four water molecules on the Pt surface. The Pt terrace (111, 110, and 100) domains exhibit high affinities for adsorbing both hydrogen atoms and water molecules.41–43 A study on the competition for adsorption on Pt (111) between hydrogen atoms and water molecules revealed that the adsorbed water molecules can destabilize both the binding behavior and the bond strength of Pt–H.44 The competition for adsorption between hydrogen atoms and water molecules on Pt is another major cause of the overpotential in the HER. This obstacle in our research is counteracted by narrowing the widths of the Pt terrace domains using SWCNT nano-fabrics as templates and supports.
Let's now focus on the role of Pt atoms on the catalyst surface. During the discharge process, electrons flow from the current collector to hydronium ions through a cathodic catalyst, primarily achieved through the Volmer step. Structural observations (Fig. 2C and ESI Fig. S10†) show that the terrace (111, 110, and 100) domains of the Pt/SWCNT hybridized nano-fabrics have a narrow width, only a few nanometers. These nanocrystalline orientations create numerous active sites on the catalyst surfaces' edges and steps. Analysis of the XPS data of the Pt/SWCNT hybridized nano-fabrics (Fig. 3B) leads to the conclusion that the active sites on the edges and steps are oxidized. These oxidized sites have semiconductive properties and can trap electrons within the metallic terrace domains. The discharge process takes place automatically through the oxidized sites once the electrons accumulate beyond the capacity of the metallic terrace domains.
We now focus on the challenges encountered when transferring electrons from current collectors to catalysts via catalyst supports. The difficulty level is mainly determined by the overall electrical resistivity encountered within the current collector, catalyst support, and catalyst. The Tafel slope is commonly used to estimate electrical resistivities involved in the Volmer step. However, the linearity of the Tafel plot is typically limited to a range of 10 to 150 mA cm−2 of the applied current densities.45 In actual hydrogen generation via electrolysis cells, the applied current densities are commonly greater than 1000 mA cm−2.46–49 Based on the EIS data (see Fig. 5D), the overall electrical resistivity involved in the HER is predominantly determined by the electrical properties of the catalyst support. SWCNT nano-fabrics exhibit excellent electrical properties, making them highly desirable as catalyst supports.
It's important to note that hydrogen spillover can hinder the formation of hydrogen molecules on the catalyst; this phenomenon must be considered, especially when high current densities are applied. Hydrogen spillover has inevitably occurred in traditional powder-form Pt/C catalysts. However, in the case of the Pt/SWCNT hybridized nano-fabric catalysts, the hydrogen spillover is highly suppressed because the surface of the catalyst support is almost completely covered by the nano-thin Pt layers.
Based on our experimental data and the DFT calculations, we propose a new model, as shown in Fig. 6, to describe the electrocatalytic functions of the active sites on edges, steps, and the terrace domains. The key concepts of this new model are summarized as follows:
(1) The initial step of the HER involves maintaining the hydronium ions within a specific range of electron discharge by creating an electric double layer (EDL). The number of hydronium ions involved in the EDL is proportionate to the surface area of the catalyst and its support.
(2) On the catalyst surface, all active sites are important but they function differently. The sites on the edges and steps are oxidized and act as atomic dams where electrons accumulate in the terrace (111, 110, and 100) domains, enabling electron discharge. The sites on the terrace (111, 110, and 100) domains are metallic and capture hydrogen atoms and water molecules, and form H–Pt and Pt/H2O, respectively.
(3) Continual H–Pt formation in the metallic terrace (111, 110, and 100) domains excludes water and leads to the hydrogen molecule formation through either Tafel's reaction: H–Pt + H–Pt′ ⇔ H2↑ + Pt + Pt′ or Heyrovsky's reaction: Pt–H−e + H+ ⇔ H2↑ + Pt.
(4) Increasing the number of active sites on edges and steps speeds up the discharge. Reducing the width of metallic terrace domains (111, 110, and 100) promotes the formation of hydrogen molecules. Both goals can be achieved using SWCNT-based nano-fabrics as the templates and supports.
(5) The durability of the Pt/SWCNT hybridized electrocatalysts is maintained through specific interactions between the nano-thin Pt layer and the highly uniform sp2-hybridized surface of the SWCNT-based nano-fabrics.
Fig. 7B presents experimental data evaluating the stability of a PEM cell with Pt/SWCNT nano-fabrics (Pt loading: 16.43 μg cm−2) as the cathodic catalyst operated at a high voltage (1.99 V) and a high current density (2.5 A). The current stability test ran for more than 300 hours, and the current densities remained almost constant at 2.5 A cm−2, indicating the superior electrocatalytic performance of the Pt/SWCNT nano-fabric catalysts even at high operating current densities. Additionally, the electrocatalytic performance of our Pt/SWCNT nano-fabric catalysts was found to be superior to that reported for the state-of-the-art nanostructured thin-film Pt catalysts.50,51
The efficiency of the PEM electrolysis cell was assessed by measuring the volume of hydrogen produced during water hydrolysis. Table 1 summarizes the cell voltage, applied current density, volume of hydrogen generation, and hydrolysis efficiency data.
Current density (A cm−2) | Voltage (V) | Hydrogen volume (mL min−1) | Theoretical hydrogen volume (mL min−1) | Electrolysis efficiency (%) |
---|---|---|---|---|
1 | 1.69 | 7.84 | 7.956 | 98.5 |
1.5 | 1.79 | 11.84 | 11.835 | 100 |
2 | 1.89 | 15.85 | 17.796 | 89.1 |
2.5 | 1.99 | 20.28 | 23.422 | 86.6 |
3 | 2.09 | 24.07 | 29.519 | 81.5 |
The electrocatalytic activity tests were carried out in a standard three-electrode system using an electrochemical analyzer (model CHI 760E, CH Instruments). A porous carbon cylinder and an Ag/AgCl electrode were used as the counter electrode and the reference electrode, respectively. The experiments were conducted in 0.5 M H2SO4 at room temperature.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04887c |
This journal is © The Royal Society of Chemistry 2024 |