Subhajit
Saha
a,
Nilankar
Diyali
a,
Sangharaj
Diyali
a,
Subhra Jyoti
Panda
b,
Mainak
Das
c,
Sobhna
Acharya
d,
Prafullya Kumar
Mudi
e,
Monika
Singh
d,
Partha Pratim
Ray
c,
Chandra Shekhar
Purohit
b and
Bhaskar
Biswas
*a
aDepartment of Chemistry, University of North Bengal, Darjeeling-734013, India. E-mail: bhaskarbiswas@nbu.ac.in; icbbiswas@gmail.com
bDepartment of Chemical Sciences, National Institute of Science Education and Research, Bhubaneswar 752050, India
cDepartment of Physics, Jadavpur University, Kolkata-700032, India
dEnergy and Environment Unit, Institute of Nano Science and Technology, Mohali 140306, India
eDepartment of Chemistry, University of Burdwan, Burdwan 713104, India
First published on 24th July 2024
Under the gravity of future socio-economic development, the viability of water electrolysis still hinges on the accessibility of stable earth-abundant electrocatalysts and net energy efficiency. This work emphasizes the design and synthesis of two newly developed cobalt(II) complexes, [Co(HL)2(NCS)2] (Comono) and [Co2(L)3(CH3OH)]ClO4 (Codi), with a (N,O)-donor ligand, HL (2-methoxy-6-(((2-methoxyphenyl)imino)methyl)phenol). The study delves into understanding their structural, morphological, magnetic, and charge transport characteristics. Moreover, the study explores the potential of these complexes in catalyzing hydrogen production through heterogeneous electrocatalysis. The X-ray crystal structure of Comono reveals the octahedral geometry of the Co(II) ion, adopting two HL units and two NCS− units. The Codi complex exhibits a doubly-phenoxo-O-bridged (μ1,1) dinuclear complex, forming a typical octahedral geometry for both the Co(II) centres in coupling with three units of L−. Temperature-dependent magnetic susceptibility measurements showed that all of the Co(II) ion in Comono shows a typical paramagnetic behaviour for high spin octahedral Co(II) ions while the Co(II) centres in Codi are coupled with doubly-phenoxo-bridges bearing weak ferromagnetic characteristics at low temperature. Electron transport properties of the Co(II) complex-mediated Schottky device address the superior carrier mobility (μ) for Codi (9.21 × 10−5) over Comono (2.02 × 10−5 m2 v−1 s−1) with respective transit times of 1.70 × 10−9 and 7.77 × 10−9 s. Additionally, electron impedance spectral analysis supports the lower electrical transport resistance of Codi relative to Comono. The heterogeneous electrocatalytic HER activity of Codi and Comono in 0.1 M KOH shows excellent electrocatalytic efficiency in terms of the various electrochemical parameters. Constant potential electrolysis, multi-cycle CVs, and post-HER analysis reveal the pre-catalytic nature of the complexes, which in turn delivers Co3O4 nanoparticles as the active catalysts for efficient hydrogen evolution.
Rationally designed coordination compounds have emerged as exciting molecular catalysts for pursuing excellent functional activities, particularly the active sites of metalloenzymes.8–11 Notably, coordination compounds offer vast possibilities for metal–ligand coordination, synergism, and facile control of the coordination environment, enabling them to be a promising platform for electrocatalytic applications.7 Ligands play a substantial role in the decoration of molecular complexes as electrocatalysts.12
Amongst the availability of different ligands, Schiff bases stand out as an important class of ligands for the judicious design of metal complexes owing to their preparative accessibility, facile tuning of donor sites, and introduction of flexibility in the ligand backbone.13–15
Additionally, the preferential choice of metal ions for the preparation of metal complexes is as crucial as the ligands. So far, many researchers have designed and developed molecular electrocatalysts based on earth-abundant 3d-metal ions in coupling with tailor-made ligands with remarkable overpotential and turnover numbers.7 However, cobalt-based molecular electrocatalysts have marked their distinctive footprints in the electrocatalytic hydrogen evolution activities due to their abundant reserves, low cost, and small energy barrier for H adsorption.16
In general, cobalt complexes have been evaluated as electrocatalysts for hydrogen evolution reaction (HER) in the non-aqueous medium using an acid source as well as in buffer media.17 Recently, Padhi and researchers developed two dinuclear cobalt(II) complexes bearing benzimidazole-derived redox-active ligands for electrocatalytic proton reduction in DMF–water (V/V, 95/5) mixture using acetic acid and elucidated the role of electronegative substituents in monitoring the catalytic activity through electron-coupled-proton transfer mechanistic routes.18 However, cobalt-based molecular electrocatalysts are prone to hydrolysis in the aqueous-alkaline medium. Typically, the molecular complexes undergo significant structural transformations under operational conditions, and therefore, the stability issue limits the use of molecular catalysts in alkaline HER.19 On the contrary, heterogeneous electrocatalytic HER-mediated by molecular complexes in alkaline conditions has marked several advantages in terms of stability, easy separation, recyclability, and high tolerance factor, where supports like carbon nanotube (CNT),20 graphene oxide (GO),21 and nickel form (NF)22 play vital roles for the deliberation of highly efficient electrocatalysts. Recently, M. Natali and researchers reported the excellent electrocatalytic activities of cobalt(II)–polypyridine complexes immobilized on single-walled carbon nanohorns (SWCNHs) with an overpotential of 0.52 V in 1 M phosphate buffer (pH = 7.4).23 Further, K. Kumar and team developed a heterogeneous electrocatalyst consisting of cobaltoxime complexes anchored on activated carbon cloth (CC) with 334 mV@10 mA cm−2 in an alkaline medium.24 Furthermore, R. G. Doria and group developed a multi-walled carbon nanotube-supported cobalt–dithiolene complex-mediated heterogeneous catalyst for hydrogen evolution studies with a remarkable TON = 50980 in pH 7 medium.25 S. Vasudevan and the group reported two mononuclear cobalt(II) complexes, [Co(OH2)2(PMBP)2], and [Co(OH2)2(PMTP)2], which exhibited overpotentials of −0.91 and −0.77 V vs. RHE@10 mA cm−2, respectively, in 1 M KOH solution.26
Nevertheless, effectual compositional and structural modification to innovate heterogeneous electrochemical characteristics in the electrocatalysts hold great promise, delineating the profound activity of the catalysts.27 The excellent catalytic performance of HER can be accredited to the upsurge of active sites, surface/interface regulation, synergistic effects, modulation in electronic structure, and the charge transfer process.28–30
In addition, the electron transport and ferromagnetic properties of a material may greatly augment the electrocatalytic activity, triggering a new avenue in the electrocatalyst design.31 Particularly, materials with large unpaired spins and a low resistance at the electrodes greatly influence the course of electrocatalytic HER.30,32 Wang and researchers demonstrated that ferromagnetic particles can afford efficient electron transfer from the glassy carbon electrode to active sites under an external vertical magnetic field, leading to an elevation in the HER performance.33,34 Therefore, determining an electrocatalyst's electron transport and magnetic properties is crucial for labelling its electrocatalytic efficiency.
In this present endeavour, we have prepared an N,O-type Schiff base ligand and synthesized two cobalt(II) complexes, Comono and Codi. The compounds were structurally and morphologically characterized with spectroscopic and analytical techniques as well as X-ray crystallography. The electron transport and temperature-dependent magnetic properties were also revealed to assess the electrocatalytic prospects of the complexes. The electrocatalytic fate of the synthetic cobalt(II) catalysts in heterogeneous alkaline HER was comprehensively studied. Constant potential electrolysis, multi-cycle CVs, and post-HER analysis suggest the in situ development of Co3O4 active catalysts for hydrogen evolution studies.
The FT IR spectra for both the cobalt complexes are shown in Fig. S1 (detailed discussion in ESI†). The UV-Vis spectrum of Comono in the acetonitrile at room temperature displays the electronic bands at 270, 333, 434 and 636 nm (Fig. S2†). The bands at ∼270 and 333 nm are assignable to the ligand-centric π–π*/n–π* transitions, while the optical band at 434 nm corresponds to the intra-ligand charge transition.35 Further, a broad d–d band from 600 to 700 nm is noticed in the UV-Vis spectrum of Comono.36 The electronic spectrum of Codi in acetonitrile shows the appearance of the electronic bands at 292, 331 and 409 nm, addressing the π–π*/n–π* electronic transition of ligand origin and the phenoxo to Co(II) charge transition octahedral Codi complex.37
X-ray structure analysis of Codi discloses that the asymmetric unit adopts three deprotonated L− units, two Co(II) centres, one coordinated methanol and perchlorate as a counter ion. In this structure, the ligational characteristics of the ligand are found to be interesting. The three ligand units show a different denticity as bi-, tri-, and tetradentate in coordination with the cobalt(II) centres. Probably, the phenolate-O-induced bridging between the Co(II) centres enables the metal ions under a crowding environment and imposes a steric hindrance among the donor sites. This results in the distal movement of the donor centres from the zone of coordination of the Co(II) centres (Fig. 1c). It is also evident that a linear geometric CH3OH is seen to coordinate with one of the Co(II) centres due to the availability of a narrow channel. A perchlorate ion is also evident for balancing the residual cationic charge for the dinuclear cobalt(II) complex. The bidentate and tetradentate ligand units use their phenoxo-units to couple the Co(II) centres and form a doubly-phenoxo-coupled Co(II) dimer. The Co(II) ions in the bridging core form a distorted planar Co1–(μ-O5)(μ-O8)2–Co2 core with an angle of 22.56° between the Co1–O5–Co2 plane and Co1–O8–Co2 plane. In the Co2 centre, O9, O8, O10, and N2 surround to compose an equatorial plane and O5 and N3 occupy the axial position. In the equatorial plane, the bond angle values for O9–Co2–N2 and O9–Co2–O10 are found to be 82.77 and 81.97°, while the angles O10–Co2–O8 and O8–Co2–N2 are noted as 92.41 and 106.88°, respectively. All the equatorial angles deviate greatly from the right angle. In contrast, the equatorial plane in Co1 comprises O5, O2, O7 and O3 atoms while O8 and N1 make the axial coordination. The equatorial bond angles, O5–Co1–O2, O5–Co1–O3, O2–Co1–O7, and O7–Co1–O3, are found to be 94.53, 87.52, 87.72 and 95.53, denoting a close occupancy to the right angles. The Co(II) complex is formed by the double μ-phenoxo bridges with a Co1⋯Co2 distance of 3.129(0) Å. In this Codi complex, the role of perchlorate ion is revealed through the investigation of supramolecular interactions. It is revealed that the perchlorate–oxygen assembles three dinuclear units through intermolecular C–H⋯O interactions, which is further supported by π⋯π interactions to develop supramolecular crystalline architecture (Fig. 1d). The non-covalent interaction parameter for Codi is given in Table S5.†
The morphology of the synthetic complexes has been judged with powder X-ray diffraction (PXRD), field emission scanning electron microscopy (FESEM), and electron dispersive X-ray (EDX) studies. The PXRD plots of Comono and Codi (Fig. 2a and b) display intensified sharp characteristic peaks up to 2θ values of 50°, which is attributed to a noticeable crystallinity in the compounds. Additionally, the experimental and theoretical fitting of the PXRD pattern matches closely with each other in both the cobalt complexes, recommending their bulk purity in the crystalline phases. The FESEM images of complexes (Fig. 2c and d) suggest that Comono and Codi exist in irregular-shaped flake-type particles. The average individual grain size of Comono and Codi is estimated at ∼0.415 μm and 0.759 μm, respectively. EDX analysis of the compounds (Fig. S3 and S4†) ensures a similar kind of elemental distribution, as evidenced by the elemental analysis and X-ray crystallography. Additionally, X-ray photoelectron spectroscopy (XPS) was carried out, which confirmed the presence of Co, C, N, S, and O elements in Comono and Co, C, N, and O in Codi. The XPS plot of both complexes is shown in the ESI (Fig. S5 and S6†).38
Fig. 2 (a) PXRD plots of Comono; (b) PXRD plots of Codi; (c) scanning electron micrograph of Comono; (d) SEM image of Codi. |
On the contrary, the room temperature χMT value for Codi was found to be 5.63 cm3 mol−1 K, which is greater than that of an isolated octahedral Co(II) ion (3.75 cm3 mol−1 K with g = 2.0 and S = 3/2) but lower than the expected magnetic value for two octahedral Co(II) centres (Fig. 3b). This observation accounts for the presence of strong spin–orbit coupling between two Co(II) centres. Upon cooling, the χMT value primarily decreases to a value of 5.33 cm3 mol−1 K till 150 K. Further lowering the temperature to 112 K, the χMT slowly decreases to reach a minimum of 5.31 cm3 mol−1 from where the χMT value starts to increase again slowly up to 75 K. The dicobalt(II) complex maintains nearly a flat region from 150 to 75 K. Upon lowering the temperature, the χMT value starts to increase sharply till 8 K, showing a χMT value of 6.05 cm3 mol−1. Further, the decrease in the temperature showed a drastic fall in the χMT value at 4.8 K.
To understand the dominance of magnetic interactions, the literature survey on the magnetic properties of the structural analogues diphenoxo-bridged dinuclear Co(II) complexes was carried out. The temperature-dependent magnetic characteristics from 300 to 8 K for the Codi complex are attributed to the origin of the spin–orbit coupling effects and ferromagnetic interaction between two cobalt centres. The final fall of the magnetic values at 4.8 K could be ascribed to the partial depopulation of excited magnetic states, zero-field splitting effects, and/or intermolecular interactions.40
It is well-observed that μ-phenoxo-bridged dinuclear Co(II) complexes in most cases are dominated by antiferromagnetic (AF) interactions as AF contributions (Jz2_xy, Jz2_z2, Jz2_x′2−y′2, Jx2−y2_x′y′, Jx2−y2_z′2, Jx2−y2_x′2−y′2), overcome the weaker ferromagnetic (F) contributions (Jxy_x′y′, Jxy_z′2, Jxy_x′2−y′2) in the Co2O2 system. Notably, four significant AF contributions are evident as Jx2−y2_x′2−y′2 AF and Jz2_x′2−y′2 AF, followed by Jx2−y2_z′2 AF and Jz2_z′2 AF, which coherently depend on the Co–O–Co angles. It is also observed that with the decrease in the Co–O–Co angle and closeness to the right angles, the AF contributions (Jx2−y2_x′2−y′2 AF, Jz2_x′2−y′2 AF, Jx2−y2_z′2 AF and Jz2_z′2 AF) decrease effectively without affecting the principal F contributions (Jxy_x′2−y′2 F and Jxy_z′2 F) in Co–O–Co.41 In the Codi complex, the Co–O–Co angles (97.38 and 99.51°) in the synthetic dicobalt(II) complex (Fig. 3b2) are the prime contributory factor for the exhibition of weak ferromagnetic interaction (Scheme 2). In addition, the magnetic coupling between the two nuclei is significantly influenced by the dihedral angle (δ) between the M–O–M plane and the phenyl plane besides the M–O–M angles (γ) (Scheme 2).42
It is documented that a large dihedral angle may reduce the extent of AF interactions. Song et al. evaluated the correlation of the Co–O–Co angle and dihedral angle in a synthesized double-phenoxo-dicobalt(II) complex.43 The researchers justified the relationship between the smaller Co–O–Co angle (γ < 99.7°) and the larger dihedral angle between the Co–O–Co plane and the phenyl plane (δ > 35.3°), with the development of preferential ferromagnetic properties in the bis(phenoxo)-bridged Co2O2 system. In the synthetic Codi complex, the Co(μ-O)2–Co core adopts the angles (γ) of 97.38(5) and 99.51(5)° for Co1–O5–Co2 and Co1–O8–Co2, respectively, while the dihedral angles (δ) between the Co1–O5–Co2 and Co1–O8–Co2 plane and the phenyl plane are 41.320(64)° and 36.383(67)°, respectively (Scheme 2). Therefore, the effect of the Co–O–Co angle and dihedral angle in the Codi complex ensures the switching of the magnetic coupling from antiferromagnetic to ferromagnetic properties. The Co–N/O bond distances (Fig. 3b1) are in good accordance with the reported bond distance values for doubly-phenoxo-bridged dicobalt(II) complex.43 A table of comparison is additionally drawn to correlate the dihedral angle (δ) and M–O–M angles (γ) of the synthetic cobalt complexes with that of the reported complexes (Table S6†), recommending the weak ferromagnetic properties in the synthetic Codi system.
To reveal the role of temperature in the electric conductivity for both devices, the I–V characteristics were monitored for the fabricated devices at various temperatures (Fig. 4a and b). The results show that the electrical conductivity of both devices increases with the increase in temperature, highlighting the typical characteristic of a semiconducting material. The conductivity at different temperature values is given in Table S7.† Further, the diode parameters for device-A and device-B of charge transport properties like series resistance (RS), ideality factor (η), and barrier height (ϕB) were evaluated by adopting the thermionic emission (TE) theory and Cheung's model following the eqn (S1)–(S7),† showing that the (RS) and ideality factor (η) for the two devices were determined from the slope and intercept of the dV/dln(I) vs. I plot (Fig. S8 and S9†). The barrier height for the two devices was extracted from the y-axis intercept of the H(I) vs. I curve (Fig. S8 and S9†). The values of the ideality factor, barrier height and Rs are tabulated in Table S8.† It was observed that the values of the ideality factor (η) deviate from unity for both the devices. However, the extent of deviation was less in device-A than in device-B, which is attributed to a higher magnitude of inhomogeneities in the barrier height of device-B. Typically, electron and hole recombination rate has a high probability in the depletion region and there may be an existence of interface states and series resistance to account for the deviation of the ideality factor from unity.44,45 Further, the I–V characteristic of the devices was explored in more detail by extracting the curves in the log scale to improve our understanding of charge transport mechanisms in the junction. In general, the current conduction mechanism is governed by the power law (I∞Vn), where n is the slope of the I vs. V curve.46 Here, the plot (Fig. 4c) exhibits two different regions under forward bias. At low bias voltage (Region-I), both the devices exhibit ohmic behaviour, i.e., (I∞V) while the other region (Region II), which exhibits a variation of current with the square of bias voltage (I∞V2), displays typical characteristics of electrical conductance (Fig. 4d). In this region, the injected carriers exhibit more numbers than the background carriers, creating a space charge field that follows the space charge limited current (SCLC) mechanism.47–49 Consequently, the Mott–Gurney (SCLC) eqn (S8)–(S10)† was employed to understand the device performance, which depends on the mobility (μ) and transit time (τ) of the devices. The mobility of device-A and device-B was found to be 9.21 × 10−5 m2 V−1 s−1 and 2.02 × 10−5 m2 V−1 s−1, respectively, as calculated from the I versus V2 plot. Additionally, the transit time of device-A and device-B was determined to be 1.70 × 10−9 s and 7.77 × 10−9 s, respectively, indicating higher electrons trapping probability of device-B. Moreover, the dielectric constant (εr) for device-A (0.7513) and device-B (0.4652) from the capacitance (C) vs. frequency (f) plot was estimated by eqn (S9)† (Fig. S10 and S11†). The value of the dielectric constant (εr) indicates the faster electrical transport in device-A. The charge transport parameters, viz., electrical mobility (μ), transit time (τ) and dielectric constant (εr), for both the devices are tabulated in Table S9.†
The impedance spectral analysis of Comono and Codi was performed at room temperature. The Nyquist plot (Fig. S12†) represents the graphical plot of the imaginary part (Z′′) vs. the real part (Z′) of the obtained impedance. The Codi, with a smaller semicircle area, exhibits a lower resistance of 5920 Ω than Comono, with a larger semicircle area corresponding to 12700 Ω resistance. This signifies the low resistance and better electron transport activities in Codi.50
The electrocatalytic fate of Comono and Codi was examined in an electrocatalytic heterogeneous hydrogen evolution reaction (HER) in 0.1 M KOH solution. For this, a homogeneous ink was prepared by sonicating the electrocatalysts with Nafion and ethanol. The details of the preparation of homogeneous ink are described in ESI.† The homogeneous ink was drop-cast on the surface of GCE to examine the heterogeneous electrocatalysis in 0.1 M KOH. To validate the molecular inertness of Comono and Codi catalysts during electrode ink preparation, we analysed the PXRD of the electrocatalyst after sonication. The PXRD pattern of the pure catalysts displayed a close match with the PXRD patterns of the recovered electrocatalysts after sonication (Fig. S15 and S16†), signifying that the molecular integrity of the catalysts was intact during the preparation of the electrode ink.
Additionally, to check the stability of the molecular complexes in 0.1 M KOH solution, both complexes were kept in the alkaline solution for 24 h. After 24 h, the electrocatalysts were recovered, recrystallized in methanol and then the PXRD were recorded (Fig. S15 and S16†). PXRD analysis suggests a close resemblance of the PXRD pattern between the pure and recovered catalysts. The PXRD comparison ensured the retention of the molecular integrity of the complexes in 0.1 M KOH after 24 h.51
After determining the molecular inertness of the complexes during the electrode's ink preparation and in 0.1 M KOH, both the electrocatalysts were immobilized on the surface of the working electrode with a similar mass loading of 0.212 mg cm−2. In cyclic voltammetry, a three-electrode setup consisting of glassy carbon as the working electrode, platinum (Pt) as the counter electrode, and Ag/AgCl as the reference electrode (saturated with KCl) was used. The linear sweep voltammetry (LSV) in the cathodic direction was recorded for both complexes in 0.1 M KOH solution with a scan rate (υ) of 5 mV s−1. The overpotential (η) values were estimated as 574 mV and 525 mV for Comono and Codi, respectively, at a current density of 10 mA cm−2 (Fig. 5a). The overpotential values for Comono and Codi were found to be comparable with those of previously reported cobalt complexes with remarkable hydrogen evolution studies.52–54
For a deep understanding of the electrochemical nature of the cobalt complexes, the Tafel slope was derived from the Tafel plot profiling η against logj by following η = blogj + a, where b and j represent the Tafel slope and current density, respectively (Fig. 5a inset). Many pioneers in this field expounded the mechanistic aspects of the alkaline HER through the Tafel slope.55 The Tafel slope values for Comono and Codi were evaluated at 115 and 146 mV dec−1, respectively, demonstrating the Volmer–Heyrovsky step as the rate-determining step in electrocatalytic hydrogen evolution (Fig. 5a inset). To reveal the kinetic nature of the electrocatalytic hydrogen evolution studies, the onset potentials were derived as 0.456 and 0.376 V vs. RHE for the Comono and Codi electrocatalysts, respectively.
The estimation of active sites in the electroactive film was pursued with directly related surface area parameters, namely, electrochemical double-layer capacitance (Cdl), electrochemical active surface area (ECSA), and roughness factor (Rf). The Cdl of Co(II) electrocatalysts was determined from the CVs at the non-faradaic region recorded at different scan rates (Fig. 5b and Fig. S17†). The current values at the middle potential of the non-faradaic region were plotted against the scan rate and the linear slope, equivalent to the Cdl (Fig. S18† and Fig. 5c).56–58 The Cdl value for Codi (3.46 μF) is higher than Comono (2.50 μF), manifesting a higher electrical energy storage capacity for the Codi catalyst. Subsequently, ECSA was also derived from eqn (S11).† The ECSA of Comono and Codi were estimated at 0.091 cm2 and 0.126 cm2, respectively, validating the higher electrochemical activity for complex Codi over Comono. Similarly, Rf was determined as 1.29 and 1.78 for Comono and Codi, respectively, by eqn (S12).†58
The intrinsic activity of the cobalt complexes was determined by calculating the charge (Q), the number of active sites (n), and turnover frequency (TOF) using eqn (S13)–(S15).† The number of active sites (n) for Comono and Codi was evaluated as 5.5610 × 10−8, and 1.0296 × 10−8 mol, while TOFs were calculated as 0.0931 s−1 and 0.5033 s−1, respectively. The electrochemical parameters, viz., overpotential (η), Tafel slope (b), TOF, Cdl, ECSA, and Rf of the Comono and Codi electrocatalyst, are tabulated in Table S10.†
The long-term electrochemical stability and durability of the synthetic cobalt(II) catalysts for hydrogen production in an alkaline water splitting were studied with constant potential electrolysis (CPE). The CPE was allowed to run for 6000 s at a constant potential at −0.49 V vs. RHE for both Comono and Codi in 0.1 M KOH. The CPE plots are displayed in Fig. S19.† The CPE plot for Codi conveys the rapid synchronization of bubble accumulation and release throughout electrocatalysis whereas the current-time plot of Comono maintains a relatively slower release and accumulation of the bubble. However, the current density for the Codi catalyst during CPE is higher than the current density for the Comono catalyst. Furthermore, the LSV of both cobalt electrocatalysts was recorded before and after CPE. For Comono, the LSV after CPE displayed an overpotential of 426 mV while for Codi, it was found to be 412 mV. After CPE, we noticed lower η values for both cobalt electrocatalysts compared to the η values of the cobalt catalysts prior to CPE (Fig. S20 and S21†). The reduction in the η values for both the electrocatalysts is attributed to the more electrocatalytically active species, manifesting the in situ generation of structurally transformed active species.19 However, the overpotential of Codi after CPE was lower with respect to Comono, signifying the transformation of more active species. The electrochemical parameter differences of both the catalysts before and after CPE are tabulated in Table S11.†
Furthermore, to gain more insight, CPEs for both catalysts were performed at a lower constant potential of −0.37 V vs. RHE with an extended time of 5 h. The CPE plot of Comono displayed a subsequent rise in the current density from −2.18 mA cm−2 to −7.12 mA cm−2 during 5 h (Fig. S22†), while for Codi, the current density increased from −2.33 mA cm−2 to −11.2 mA cm−2 (Fig. 5d). The gradual change in the current density during CPE suggests that parallel faradaic processes are occurring during the electrocatalytic hydrogen evolution. The faradaic Efficiency (FE) calculation for the first 1 h of CPE was found to be 35.7% and 31.2% for Comono and Codi, respectively, signifying some concurrent electrochemical processes during CPE. However, after one hour of CPE, Comono and Codi displayed a steady current density with better FE of 89.8% and 93.6%, respectively. This confirms that the electrocatalysts experience significant dynamic irreversible activation within the first hour of CPE. In addition, multi-cycle CV (500 times) was performed for both complexes at 1000 mV s−1 scan rate in sweeping between 0.96 and −0.84 (V vs. RHE) (Fig. S23 and S24†). A significant current elevation was evident for both the catalysts. The multi-cycle CV analysis also consolidates the formation of more active catalytic species under the electrochemical conditions. However, the elevation in the current change is higher in magnitude for Codi than that for Comono.
Upon confirming the concurrent generation of structurally derived active species, XPS analysis was conducted to investigate the nature of the recovered active species after HER. The high-resolution Co 2p XPS spectra for post-electrochemical species are shown in Fig. S25.† XPS analysis reveal a significant difference in the binding energy value of the pure complex and post-electrochemical species corresponding to Co 2p3/2.38,59 For Comono, the peak at 781.08 eV was shifted to 780.5 eV after HER. Similarly, the peak at 781.3 eV of Codi shifted to 780.7 after HER (Fig. S25a and S25b†). Furthermore, the nature of post-electrochemical active species was elucidated by PXRD for Comono and Codi. Post-CPE Comono and post-CPE Codi species displayed a high similarity in the PXRD patterns (Fig. S26†). In both the cases, the characteristic peaks at 31.3, 36.9, 38.6, 44.9, 55.7 and 59.5 were evident, corresponding to the Co3O4 (JCPDS: 00-042-1467) (Fig. S26†). The SEM images of the post-CPE cobalt species were also recorded to reveal the changes in the morphology between the cobalt complexes and the post-CPE cobalt species (Fig. S27†). The distinct change in the morphology and the particle size suggests the in situ transformations of the molecular complexes into cobalt oxide electrocatalysts.
This observation reveals that Comono and Codi pre-catalysts under electrochemical conditions undergo irreversible activation, leading to Co3O4 active catalysts. The lability associated with the cobalt complexes may be accounted for in light of the X-ray structures of the complexes. It is observed that the utilization of lower denticity (two-donors) by the synthetic polydentate (four-donors) chelating ligand in the formation of the Comono complex imposes relatively higher lability to the complex, which is attributed to the prompt hydrolysis tendency of Comono under electrochemical conditions. In contrast, the chelating ligand behaves as a multidentate ligand (4/3/2-donor) in the formation of the dinuclear cobalt complex, where the chelating-cum-bridging coordination modes by the ligand transport into its active species in the electrochemical condition. However, Codi-derived active Co3O4 species in the electrochemical conditions showed better activity compared to the Comono-derived Co3O4 species owing to the differences in the particle size and morphology.
Nonetheless, the heterogeneous nature of both the catalysts was confirmed by the wash test.60 In the wash test, the catalysts were initially subjected to CPE at a constant potential of −0.49 V vs. RHE for 6000 s. Afterwards, the working electrode was replaced by a fresh GC electrode in the same solution and CPE was taken under a similar electrochemical condition. The CPE of fresh GCE was found to be silent for both the cobalt catalysts (Fig. S28 and S29†). This observation confirmed that the solution did not contain any active electrocatalysts as well as no dissolution of Pt-electrode occurred in the electrocatalytic HER, which ensures the heterogeneous nature of the electrocatalysis.
Footnote |
† Electronic supplementary information (ESI) available: Supplementary crystallographic data for Comono and Codi. CCDC 2290443 and 2290444. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt01358a |
This journal is © The Royal Society of Chemistry 2024 |