Efrat Shawat
Avraham‡
ae,
Bibhudatta
Malik‡
ae,
Alina
Yarmolenko‡
ae,
Rajashree
Konar
abe,
Sergei
Remennik
c,
Gili Cohen
Taguri
a,
Sandro
Zorzi
d,
Elti
Cattaruzza
d,
Michael Yakov
Hubner
a and
Gilbert Daniel
Nessim
*ae
aDepartment of Chemistry and Institute of Nanotechnology and Advanced Materials (BINA), Bar-Ilan University, Ramat-Gan, 5290002, Israel. E-mail: gdnessim@biu.ac.il
bAtomic Structure-Composition of Materials, INL – International Iberian Nanotechnology Laboratory, Avenida Mestre José Veiga s/n4715-330, Braga, Portugal
cCenter for Nanoscience and Nanotechnology, Hebrew University, Jerusalem 9190401, Israel
dDepartment of Molecular Sciences and Nanosystems, “Università Ca' Foscari Venezia”, Via Torino 155, 30172 Venezia Mestre, Italy
eIsrael National Institute of Energy Storage (INIES), Ramat-Gan 52900, Israel
First published on 13th August 2024
Developing high performance catalysts for electrochemical water splitting is critical for an efficient and sustainable route to hydrogen production. For this, single-atom catalysts (SACs) are the best candidates, as they offer the highest atom efficiency. However, current methods to produce SACs involve a complex synthesis, often requiring multiple lengthy and expensive steps and yielding an insufficient density of single atoms. Here, we report a one-step chemical vapor deposition (CVD) synthesis to produce free-standing (FS) electrodes with Ni SACs on a matrix of sulfur-doped carbon nanofibers (CNFs), referred to as SACs@nanocarbon. The mechanism is based on a temperature-controlled delamination of thin films, with Au in contact with a SiO2 substrate, leading to the nucleation and growth of SACs@nanocarbon. Advanced characterization methods indicate the presence of Ni and Au single atoms and larger gold aggregates on the CNF matrix surface. These non-platinum group metal (non-PGM) electrodes showed exceptional performance for the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER). They performed for over 20000 cycles with negligible change in overpotential at higher currents, with low onset overpotentials of 305 mV at 10 mA cm−2 for the OER and 40 mV at 17 mA cm−2 for the HER. The overpotential decreased to 195 mV at a current density of 100 mA cm−2. Remarkably, the electrode performance improved over cycling, while gold was dissolving in the electrolyte. This novel synthesis yielding SACs@nanocarbon could pave the way for the development of non-PGM, high performance electrodes for many other electrocatalytic applications. Additionally, the new paradigm of temperature-controlled delamination of thin films could be used to synthesize new materials.
The HER and OER performances of a wide variety of noble and non-noble metal SACs were examined in the literature.12 Ni-, Co-, and Fe-based SACs on heteroatom-doped carbon systems showed highly efficient OER electrocatalysis in alkaline medium. Ni-doped carbon exhibited the best performance (followed by Co and Fe).13,14 Unfortunately, producing high-density SACs presents many challenges as single atoms tend to aggregate into metal clusters to lower their surface energy during the synthesis.15 Various strategies such as pyrolysis, chemical vapor deposition (CVD), atomic layer deposition, wet chemistry, atom trapping, and photochemical methods have been employed to limit this behavior.16–19 However, these methods present clear drawbacks: pyrolysis produces carbon-based supports with uncontrollable and unpredictable structures due to the degradation of metal–organic complexes. Wet chemistry methods suffer from low SAC loading and a high possibility of aggregation of metal nanoparticles.20 Atom trapping is a complicated process requiring high temperature, mobile atoms, and supports that can trap mobile species.21 In most cases, the fabrication of electrodes requires many complex and expensive steps (e.g., photolithography). Besides the above, the main issue is the low metal loadings that limit the commercial use of SACs.14,22
For practical applications, we need long-term stability and robustness of electrocatalysts, in addition to their electrocatalytic activity (e.g., overpotential, kinetics, active surface area, faradaic efficiency, and turnover frequency). Electrodes based on powders need binders (e.g., conducting polymers, mainly Nafion) for immobilizing the catalytic material on the current collector substrate. The issue with binders is that catalysts often peel off from the current collector during electrocatalysis, thus inhibiting long-term electrolysis and reducing current density.23,24 This issue can be resolved by using free-standing (FS) electrodes, made of one material without any binder. Furthermore, bi-functional catalysts for the OER/HER can lower the production cost of electrodes for total water splitting, but obtaining high performance catalysts remains a challenge.25
Is it possible to synthesize high performance, non-PGM, free-standing, SA-based, bifunctional electrodes for water splitting using a simple and economical process?
Here we demonstrate a one-step process using CVD to produce free-standing electrodes, offers several advantages, both in terms of process simplicity and cost, and also in terms of good electrode performance and durability.
More specifically, we use CVD to synthesize bi-functional, self-standing electrodes made of SACs on a sulfur-doped nanocarbon matrix (SACs@nanocarbon) based on thermally controlled delamination of thin films of Ni and Au. We produced SACs@nanocarbon using a one-step chemical vapor deposition (CVD) synthesis that included four consecutive processes: (1) delamination of thin films from the substrate, (2) nucleation and growth of a nanocarbon mat, (3) formation of SACs, and (4) sulfur doping of the CNFs. This process is simple (one-step), economical, and industrially scalable (CVD).
The advantage of the electrodes produced is that they are bifunctional catalysts (OER and HER) and exhibit low overpotentials and extremely high cyclability (>20000 cycles). The high conductivity of the sulfur-doped nanofiber carbon mat and its near super-hydrophobicity increase its electrical conductivity and facilitate the homogeneous diffusion of electro-active species and the weak interaction with water molecules further boosts the catalytic stability. The formed carbons are crystalline in nature, which makes them suitable for use as electrodes in water electrolysis and their extended conductive network accommodates single atoms (100% atom efficiency for catalytic reactions) to catalyze the HER and OER efficiently.
We placed the sample in an alumina boat at the exit of the downstream furnace outside the heated zone. First, we purged the reactor with 100 [sccm] of He gas (99.9999%, Gas Technologies) at room temperature for 15 min until the furnace reached the desired temperature. We then introduced 100 [sccm] of He, 200 [sccm] of H2 and 200 [sccm] of C2H4 to delaminate the thin film stack and to nucleate and grow the nanocarbon mat. Using the “fast-heat” technique described in our previous publication,26 the sample was kept at room temperature still positioned outside the exit of the second furnace and entered the heated zone only when the furnace reached the desired temperature. We thus heated the sample without exposing it to a temperature ramp. When the furnace reached 800 °C, the boat with the sample was introduced into the heated zone by sliding the quartz tube and heating for the required duration (30 s to 2 h). At the end of the synthesis, the tube was pushed out to position the boat containing the wafer outside the heating zone in order to cool it under a He flow. We labeled the delaminated mats obtained after reacting for 10 min to 2 h as C-10 (10 min), C-30 (30 min), C-60 (1 h) and C-120 (2 h) (photographs in Fig. S2†).
We purged the system using 100 [sccm] of He (99.9999%) for 15 min while keeping the two boats at room temperature (outside the heated zones) until the furnace reached equilibrium at the desired temperatures in the two zones to fill the reactor with helium, thus removing all air. To grow the nanocarbon with the SACs, we shifted the quartz tube to put the second boat inside the downstream furnace for 30 s to 2 h while flowing 100 [sccm] of He, 200 [sccm] of H2 and 200 [sccm] of C2H4 gases. After that, using an external magnet,27 we introduced the boat with the sulfur powder (1 g, Alfa Aesar, 99.5%, 325 mesh) in the upstream furnace for 60 min under a flow of 100 [sccm] of He to transport the vaporized sulfur gas to the downstream furnace to react with the sample.
At the end of the synthesis, the boat with the sulfur was pulled out of the heated zone using the magnet and the quartz tube was pushed out to position again the boat containing the sample outside the heated zone to cool under a flow of helium. The sulfur-doped delaminated mats obtained after reacting for 10, 30, 60 and 120 min are labeled S-C-10, S-C-30, S-C-60 and S-C-120 (photographs in Fig. S2†).
In a similar way, we took 10 mg of both RuO2 and Pt/C and added them to a mixture of 500 μL of DI water, 400 μL of absolute ethanol 100 μL of 5% Nafion solution and sonicated for 30 minutes. 100 μL of each catalyst ink was dropcast on previously cleaned carbon papers (1 cm2). Finally, the electrodes were dried and used for electrochemical studies.
Electrochemical impedance spectroscopy (EIS) is useful to understand charge transfer mechanisms during the electro-oxidation of water to O2. Here, we varied the frequency in the range of 100 kHz to 100 mHz (a small perturbation of 10 mV). The potential scale as per the reversible hydrogen electrode (RHE) was calibrated to ERHE = EHg/HgO + 0.93 V and thus we calculated the overpotential as η = ERHE − 1.23 V. We carried out the OER and HER studies in 1 M KOH (pH 14) under ambient conditions with saturation of the electrolyte using argon gas. While testing our samples, we observed that only those synthesized for 1 or 2 h exhibited sufficient mechanical strength to be used as freestanding (FS) electrodes.
After synthesizing free-standing electrodes with sufficient mechanical strength for electrocatalysis and superior OER performance, we stopped using powders. These binder-free self-standing electrodes do not exhibit the issues mentioned above for powder-based/binder electrodes. We analyzed the OER and HER performance of the FS electrodes by electrochemical testing in a typical three-electrode setup with a PTFE replaceable holder as the current collector for the working platinum electrode (see the optical image in Fig. S10†). This one-step synthesis, without post-processing or coating, is carried out to avoid organic binders. The electroactive species with the conductive substrates exhibit good electronic conductivity and mechanical integrity as catalysts for long-term cycling.
In the ESI,† we provide additional details on the characterization methods we used, such as XRD, XPS, Raman, etc., as described in the results section.
Raman spectroscopy measurements of S-C-120 show a G band peak at ∼1580 cm−1 and a 2D band peak at ∼2700 cm−1 (Fig. S5†). The ratio D/G = 1.08 indicates a graphitic structure, where the D band indicates a defect in the graphitic structure, which arises from out-of-plane vibration, while the G band originates from the in-plane vibration of the C–C bond from sp2 hybridization.
We characterized the C-120 (Fig. S6a†) and S-C-120 (Fig.S6b†) samples using aberration-corrected HRTEM with high-angle annular dark-field imaging (HAADF-STEM) to achieve atomic resolution. We identify the bright dots circled in yellow in Fig. 2b as single or very few atoms that we will call single atom catalysts (SACs). We performed extensive electron dispersive spectroscopy (EDS) and subsequent mapping of the different elements on the S-C-120 sample (Fig. S7†). This mapping shows that the dots are made of Ni, Au, or possibly both (see Fig. S7c and e†) and that carbon and sulfur are uniformly distributed (Fig. S10d and f†). The dimensions of the SACs, in the range of Angstroms, are in agreement with previous reports of Ni SACs (and Fe) on different supports.33 These data support the claims regarding the formation of well defined Ni SACs on the carbon matrix. Further, the HAADF imaging of the S-C-120 sample also reveals the presence of Au particles of around 200 nm in size (Fig. S8†). We characterized a cross-section (lamella) of the sample using a FIB to see how Ni penetrates the carbon material (Fig. S3†). Over a depth of around 10 microns, we observed that the concentration of nickel is at its lowest on the surface of the nanocarbon mat and increases with depth. This observation is consistent with XPS measurements (discussed later), where we only detected Ni at the bottom of the nanocarbon (i.e., the part of the electrode in contact with the substrate before delamination), thus indicating that the amount of Ni on the surface of the nanocarbon increases with depth (Fig. S7†). We quantified the atomic content of nickel using inductively coupled plasma-mass spectrometry (ICP-MS) and it revealed that S-C-120 contains 2.6% Ni in it.
Fig. 2c shows the XRD pattern of C-120, S-C-120 and post-OER (S-C-120) samples. Each material showed four main broad peaks present at about 25.8°, 43.3°, 53.8° and 78.3° that correspond to the (002), (100), (004) and (110) planes of graphite, respectively. These strongly broadened peaks typically arise from their turbostratic character. As shown in Table 1, the (002) interlayer spacing for all samples is close to the characteristic interlayer distance (0.344 nm) in turbostratic stacking. The diffraction peaks at 2θ of 38.1°, 44.3° and 64.5° correspond to the (111), (200) and (220) planes of Au (PDF #00-004-0784), respectively. After the OER, we observed the formation of the K(HCO3) phase (PDF# 01-086-0912) coming from the electrolyte. The XRD pattern suggested that no Ni phase was detected, further hinting that Ni was present only as single atoms. Structural parameters determined by quantitative analysis of the (002) peak are given in Table 1. The values of the interlayer spacing for d002 are calculated using Bragg's law as d002 = nλ/2sinθ, and the crystallite height (stack height) (Lc) is estimated via Scherrer's equation Lc = 0.94λ/B
cos
θ, where B is the FWHM and θ is the Bragg angle of the (002) band for Lc.34 The sulfurization step clearly improved the graphitic order through the decrease of d002 and the increase of Lc for both S-C-120 and post-S-C-120. Lower values of d002 indicate fewer defects in the graphitic layers, leading to better packing of the carbon nanofibers.35
Sample | d(002) [nm] | L c [nm] |
---|---|---|
C-120 | 0.348 | 2.7 |
S-C-120 | 0.344 | 3.5 |
Post-OER, S-C-120 | 0.343 | 3.6 |
Using XPS, we probed the chemical states and bonding environments of the elements present in the doped samples. The C 1s spectrum in Fig. 6a reveals the presence of sp2 hybridized carbon (peak position at 284.7 eV) and C–S bonding and π–π* transition as the deconvoluted peaks at 285.4 and 290.5 eV, respectively (Fig. S9a†).36 The high-resolution spectrum of S 2p can be divided into S 2p3/2 and S 2p1/2 at 164.2 and 165.5 eV, respectively, attributed to the C–S linkage. An additional peak is observed at 168.5 eV due to SOx bonding (Fig. S9b†).37,38 The core-level spectrum of Ni 2p3/2 is deconvoluted into three peaks positioned at 856.2, 861.5 and 862.8 eV, which can be assigned to high valence Ni (Ni3+) and satellite peaks (Fig. S9c†).39 Fig. S9d† illustrates the XPS spectrum of Au 4f, where two well-defined peaks at 84.1 and 88.1 eV correspond to Au 4f7/2 and Au 4f5/2, respectively, suggesting the existence of metallic Au.40
We also performed contact angle measurements (Fig. S10†). The contact angle of the non-doped sample was 128.7° compared to 143.2° for the sample doped with sulfur, indicating that the doped sample is more hydrophobic (almost super-hydrophobic), which will be important for explaining the electrocatalytic performance.
The complex mechanisms leading to the formation of the SACs are difficult to precisely define without a dedicated in situ reactor that allows real-time characterization. This is compounded by several parameters playing a role, such as reaction time, temperature, amount of S doping, and the nature of the cooling procedure in this single-step CVD reaction. However, we can propose a plausible mechanism based on the extensive characterization performed.
The outstanding questions are mainly:
(1) Why do Ni and Au form SACs on the surface of the nanocarbon?
(2) Besides helping the delamination at high temperature, do Ni and Au contribute to the nucleation and growth of the carbon matrix?
(3) What is the role of sulfur? (As will be explained later, it significantly improved the electrocatalytic performance).
It is easy to understand that the Ni/Au thin film stack delaminates from the substrate owing to the weak adhesion of gold to Si/SiO2. We can also hypothesize that the stack then fragmented into sub-micron particles that, in the presence of ethylene at high temperature, started nucleating and growing carbon nanofibers (Fig. 1). This is a plausible hypothesis given that Ni is a good catalyst for carbon nanotubes and nanofibers.41,42 It is validated by our previous research, where we studied a delaminated layer of Ni/Pd weakly adhered to the silicon wafer with a very thin layer of Ti (as Pd does not adhere at all to Si/SiO2), where, based on extensive characterization, we formulated a mechanism based on the fragmentation of the thin film into Ni/Pd alloy particles that led to nucleation and bidirectional growth of the nanofibers.31 However, in that work, no metal atoms were present on the surface of the nanofibers, and the sample did not exhibit any electrocatalytic activity.
Here, we used Au as a weak adhesion layer and observed the formation of CNFs (as for Ni/Pd) but also found metal atoms on the surface of the carbon. We may explain this difference by observing and comparing the bulk phase diagrams of Ni–Pd and Ni–Au.§ The Ni–Pd phase diagram shows complete miscibility below 1237 °C for any composition. However, the Ni–Au phase diagram shows a synclinal miscibility gap with a peak at 807 °C at 72 at% Ni. From this, we can hypothesize the formation of two phases of the Ni–Au alloy during the thermal process with one phase slightly richer in Ni. Given that Ni is a known catalyst for CNF growth, we can speculate that one of the two phases is responsible for the nucleation and growth of the CNFs after fragmentation (similar to our previous work),31 while the second phase “dewets” on the surface to form Ni and Au SACs.
The excellent electrocatalytic performance confirms that Ni SACs play a key role. However, regarding the Au SACs, we find gold in the solution during OER cycling (starting after 250 cycles and observed until 50000 cycles), indicating that Au atoms on top of the carbon leach out to the electrolyte solution. Interestingly, the performance of the electrode increased during the leaching of the gold. Note that we also found larger gold chunks on the nanocarbon matrix that may also detach and dissolve into the electrolyte solution during electrocatalysis (Fig. S8†).
Let us now analyze the role of sulfur. Doping carbon nanomaterials is a practical and powerful approach to improve their properties, in particular their electrical conductivity.43 Heteroatom-doped carbon nanotubes (CNTs) are studied in many fields, such as biotechnology, catalysis, optics, photovoltaics, supercapacitors, and energy storage. Based on the unique electronic properties and high surface area of CNTs, as well as the similar electronegativity of sulfur and carbon, Li et al.44 prepared a novel electrocatalyst for the oxygen reduction reaction (ORR) by directly annealing oxidized CNTs and p-benzene dithiol in nitrogen. Previous studies demonstrated both experimentally and theoretically that heteroatom-doped CNTs can be regarded as a promising electrocatalyst for the OER.38
In our research, we also chose to dope our material with sulfur. Previous reports reveal that the inclusion of sulfur atoms in the carbon lattice may regulate the charge redistribution by tuning the electronic structure of the catalytic center, where adsorbates such as OOH*, O* and OH* prefer to adsorb.45 We evaluated the samples with and without doping and observed that the electrocatalytic performance of the S-doped electrode was always far superior. We believe that the increased electrical conductivity and hydrophobicity due to sulfur doping are responsible for the improved electrochemical performance. Earlier research shows that the doping of S atoms in the carbon lattice has a great impact on the electronic structure and spin density and generates a significant number of active sites on carbon atoms, thus exhibiting improved electrocatalytic activities compared to pristine carbon samples.46
We then performed LSV of the sulfur-doped ground samples and observed a different trend (Fig. S11b†). The S-C-10, S-C-30, S-C-60 and S-C-120 powders reached the same current density of 10 mA cm−2 at a respective overpotential of 420, 470, 480 and 400 mV, clearly indicating a significant effect of sulfur on OER activity. Among these samples, S-C-120 outperformed the other synthesized materials in terms of lower overpotential, with the activity trend being:
S-C-120 > S-C-10 > S-C-30 > S-C-60. |
For all powder samples, we analyzed the kinetics of the synthesized delaminated electrodes using the Tafel equation [η = a + b × logj] with a = constant, b = 2.3RT/αF, and j = current density. We carried out LSV with a rotating disk electrode (RDE) at 1600 rpm at a scan rate of 5 mV s−1. The ohmic drop corrected Tafel slopes were 51, 55, 61 and 58 mV dec−1 for the non-doped C-10, C-30, C-60 and C-120, respectively (Fig. S11c†). C-10 showed faster kinetics while oxidizing water. The sulfur-doped S-C-10, S-C-30, S-C-60 and S-C-120 exhibited slopes of 57, 54, 75 and 50 mV dec−1 with the smaller Tafel slope of S-C-120 validating its improved OER kinetics (Fig. S11d†).
![]() | ||
Fig. 3 CVs of (a) C-120_FS and (b) S-C-120_FS and (c) linear sweep voltammetry (LSV) and (d) Tafel plots of the two delaminated FS electrodes. |
As previously mentioned, the performance of the sulfur-doped FS electrode was much better compared to the pristine electrode, likely due to the improved electronic conductivity due to S-doping (from impedance measurements). Furthermore, we noticed the dissolution of gold into the electrolyte solution (see Fig. S10 and S20†), consistent with the report that Au dissolution starts at 1.2 V vs. RHE in an alkaline solution.48 After 500 cycles, we noticed a gradual color change of the electrolyte from shiny yellow to dark red (optical images are shown in Fig. S10†) as Au atoms or nanoparticles detach from the carbon. We can speculate that this leads to defects on the carbon matrix that further activate the surface to better catalyze the OER.
LSV that was carried out at 5 mV s−1 suggests that S-C-120_FS delivered a benchmarking current density of 10 mA cm−2 at an overpotential (η) of 300 mV and showed much higher OER activity than C-120_FS, which exhibited η @10 mA cm−2 = 350 mV (Fig. 3c). S-C-120_FS and C-120_FS delivered a current density of 50 mA cm−2 at the corresponding η of 380 and 420 mV. S-C-120_FS and C-120_FS reached a j of 100 mA cm−2 at the respective overpotentials of 430 and 470 mV. Further, the OER activity of S-C-120_FS and C-120_FS outperformed the OER activity of state-of-the-art RuO2. RuO2 reached a current density of 10 mA cm−2 at an overpotential of 340 mV. We studied the reaction kinetics using their corresponding Tafel plots extracted from the LSV curves, as shown in Fig. 3d. S-C-120_FS exhibited a smaller Tafel slope of 65 mV dec−1, while C-120_FS exhibited a slightly higher slope of 68 mV dec−1. We can speculate that the higher slopes of the FS samples compared to the same samples from powders could be due to the formation of large gas bubbles attached to the electrode surface.
The electrochemically active surface area (ECSA) is a useful descriptor for an electrocatalyst. The ECSA can be calculated as ECSA = Cdl/Cs, where Cs is the specific capacitance of the electrode (the Cs value can be 0.04 mF cm−2 in 1 M KOH as per reports49) and Cdl denotes the double-layer capacitance, which is assessed from CV measurements at different sweep rates carried out in the non-faradaic regions (no interfacial charge transfer), as depicted in Fig. S12.† C-120_FS exhibited a Cdl of 98 mF cm−2, whereas S-C-120_FS exhibits a much higher Cdl of 257 mF cm−2, as shown in Fig. 4a and their corresponding ECSAs are 2450 and 6425 cm2. Thus, S-doping not only lowers the overpotential, but also significantly enhances the ECSA.
We assessed the stability of S-C-120_FS using potentiostatic chronoamperometry. We applied a potential of 1.61 V vs. RHE (corresponding j = 50 mA cm−2) to observe the change in current density over time. As we can see in Fig. 4b, the current became steady nearly after 2.5 h to deliver 50 mA cm−2, showing excellent stability up to 24 hours (with a slight degradation). After 50 h of constant electrolysis, the current density retained is 68% of its initial value. The interruption and reduction of j occurred after 24 hours and this could be attributed to the accumulation of the formed O2 gas bubbles at the surface of the electrode or to the constant Au release into the electrolyte, which eventually reduced the conductivity of 1 M KOH. It is reported that the design of superhydrophobic surfaces could assist in bubble detachment, thus making the electrode more robust.50 Contact angle measurements revealed that S-C-120_FS was almost superhydrophobic (143.2°; Fig. S17†), which could be the reason for its remarkable stability. The OER durability of the electrode is shown by LSV after 5000 CV cycles, which exhibited negligible loss of potential (Fig. S16†). We also tested the OER durability of the electrode for about 20000 cycles and observed a large change in potential (Δη at 25 mA cm−2 = 80 mV) in the same electrolyte; however, when we switched to a fresh electrolyte, we observed a very pronounced activity with negligible loss in potential (Δη at 25 mA cm−2 = 10 mV), better than what we observed after 5000 cycles.
In the initial cycles, S-C-120_FS exhibited high charge transfer resistances, given that the diameter of the semicircle in the Nyquist plot is quite large, leading to poor conductivity and slower OER kinetics (Fig. 4c). However, after 2000 cycles, S-C-120_FS became significantly active towards faradaic reactions, as shown in the Nyquist plot (Fig. 4d). The distance between the high frequency and low frequency regions is greatly reduced (a smaller arc of semicircles), indicating better charge transfer resistance and improved electronic conductivity. The two types of semicircles observed for S-C-120_FS can be attributed to the presence of two time constants (see the Bode plot in Fig. S17† and the equivalent circuit with fitting parameters in Fig. S18†).
EIS spectra after 20 k cycles revealed that S-C-120_FS showed a much smaller semicircle than the initial cycles of S-C-120_FS and C-120_FS, thus indicating better conductivity after 20 k cycles (Fig. 5c). We observed two atypical incomplete semicircles in these three cases: one small semicircle at higher frequency representing bulk electrolyte resistance and a large semicircle formed at lower frequency representing the charge transfer resistance owing to H2 evolution. We performed chronoamperometry to study the stability of the electrode after 20 k cycles (Fig. 5d). The electrode showed good HER stability over 100 h of electrolysis at a high current density of 100 mA cm−2 (η of 195 mV), with the j retained at 80%, with the current drop possibly caused by the accumulation of gaseous bubbles. Overall, the results imply that the diffusion of the electrolyte into the electrodes (for wetting) took a longer time to activate the electrodes (∼2500 CV cycles at 20 mVs−1) and to allow the reactant species (H2O) to reach the catalytically active sites.
The survey spectrum after the OER showed the presence of O, C, S and Ni (see the zoomed-in spectra of Fig. S19†) and additionally K (K 2s at 378.2 eV, K 2p at 293.4 and 296.2 eV, K 3 s at 33.8 eV and K 3p at 17.6 eV) was also present (Fig. S19†). The C 1s spectrum can be fitted with four peaks at 284.5 eV (adventitious carbon), 284.7 eV (carbon in the matrix of the material), 285.6 eV (carbon bound to S) and 288.15 eV (CO). The S 2p doublet at 164 eV is compatible with S bound to C (thiophene type), while the S 2p doublet at 168.3 eV can be attributed to SOx species. Ni likely exists in an +3 state, as evidenced by the presence of a 3p3/2 characteristic peak at 69.05 eV.52,53 The O 1s spectrum can be fitted to three peaks at 531.4, 533.15 and 535.15 eV due to M–O, M–OH and M–H2O bonding, respectively.
In the survey spectrum, after the HER we detected O, C, S and Ni (see zoomed-in spectra, Fig. S20†). The C 1s spectrum can be fitted with three peaks at 284.45 eV (adventitious carbon), 284.7 eV (carbon in the matrix of the material) and 285.5 eV (carbon bound to S). The S 2p doublet at 163.75 eV is compatible with S bound to C, while the S 2p doublet at 168 eV can be attributed to SOx species. Ni in this case exists in a +2 oxidation state where the observed peak is positioned at 68.25 eV.53 The O 1s spectrum can be fitted with two peaks at 531.7 and 532.2 eV, which can be attributed to metal–oxygen or carbon–oxygen bonding.
Many physicochemical effects are at are involved in delivering this material, including thermal delamination of thin metal films via a weak adhesion layer, the nucleation and growth of carbon nanofibers, subsequent sulfur doping, the formation of Ni and Au SACs, and changes in hydrophobicity during cycling. We believe that this new paradigm of thermal delamination of thin films to grow catalyst atoms on a doped nanocarbon matrix could open new research avenues to produce high-performance, low-cost electrodes for various electrochemical applications.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04701j |
‡ These authors have made equal contributions to this work. |
§ Although our metallic films are tens of nanometers thick, which could lead to melting point depression or other phenomena that differ from the bulk, we will use as a first approximation the bulk phase diagrams to hypothesize the growth mechanisms. |
This journal is © The Royal Society of Chemistry 2024 |