Aldona Balčiūnaitė†
a,
Kush K. Upadhyay†b,
Kristina Radinovićc,
Diogo M. F. Santosd,
M. F. Montemorb and
Biljana Šljukić*cd
aDepartment of Catalysis, Center for Physical Sciences and Technology, Saulėtekio ave. 3, Vilnius LT-10257, Lithuania
bCentro de Química Estrutural-CQE, Institute of Molecular Sciences, Departamento de Engenharia Química, Instituto Superior Técnico, Universidade de Lisboa, 1049-001 Lisbon, Portugal
cUniversity of Belgrade, Faculty of Physical Chemistry, Studentski trg 12-16, 11158 Belgrade, Serbia. E-mail: biljana.paunkovic@tecnico.ulisboa.pt
dCenter of Physics and Engineering of Advanced Materials, Laboratory for Physics of Materials and Emerging Technologies, Chemical Engineering Department, Instituto Superior Técnico, Universidade de Lisboa, 1049-001 Lisbon, Portugal
First published on 30th March 2022
β-Ni(OH)2 nanoplatelets are prepared by a hydrothermal procedure and characterized by scanning and transmission electron microscopy, X-ray diffraction analysis, Raman spectroscopy, and X-ray photoelectron spectroscopy. The material is demonstrated to be an efficient electrocatalyst for oxygen reduction, oxygen evolution, and hydrogen evolution reactions in alkaline media. β-Ni(OH)2 shows an overpotential of 498 mV to reach 10 mA cm−2 towards oxygen evolution, with a Tafel slope of 149 mV dec−1 (decreasing to 99 mV dec−1 at 75 °C), along with superior stability as evidenced by chronoamperometric measurements. Similarly, a low overpotential of −333 mV to reach 10 mA cm−2 (decreasing to only −65 mV at 75 °C) toward hydrogen evolution with a Tafel slope of −230 mV dec−1 is observed. Finally, β-Ni(OH)2 exhibits a noteworthy performance for the ORR, as evidenced by a low Tafel slope of −78 mV dec−1 and a number of exchanged electrons of 4.01 (indicating direct 4e−-oxygen reduction), whereas there are only a few previous reports on modest ORR activity of pure Ni(OH)2.
For complete water splitting to produce simultaneously hydrogen and oxygen gases, cathodic (hydrogen evolution reaction, HER) and anodic (OER) reactions should be efficient in the same electrolytic solution.18–20 The disparity of the conditions under which the electrocatalyst effectively operates is responsible for a decrease in the efficiency of the water-splitting process.21–23 Materials with electrocatalytic activity toward HER typically show better performance in acidic electrolytes, while those with activity toward OER generally show good performance in alkaline electrolytes. Building bifunctional electrocatalysts with high efficiency for both hydrogen and oxygen evolution, as well as stability in the same electrolytic solution, has been marked to be of utter importance. As mentioned above, the electrocatalyst's structure and morphology should be designed to display large electrochemically active surface area, fast charge transfer, and gas transport. These properties are necessary to achieve high electrocatalytic activity, decrease the overpotential of the evolution reactions, and enhance the reaction kinetics. Overall, this leads, finally, to a significant increase in the water-splitting process efficiency.23,24
Besides the two reactions (OER and HER) involved in water splitting, the oxygen reduction reaction (ORR) is vital in electrochemical energy storage/conversion devices, such as fuel cells and metal–air batteries.25,26 The main limiting factor in the energy conversion efficiency of these devices is the sluggish ORR kinetics since ORR requires high overpotential and has a complex mechanism involving several steps.27,28 ORR in alkaline media (such as in alkaline fuel cells) has faster kinetics, allowing the use of non-platinum (Pt) electrocatalysts in these cells.29 Still, there are very few reports on Ni(OH)2 activity for ORR; these previous studies typically involve supported Ni(OH)2, transition metals (e.g., Mn) coupled with Ni(OH)2,30 or Ni(OH)2 supported on graphene oxide.31,32 The activity of pure/unsupported Ni(OH)2 electrocatalysts reported in those studies was rather modest.31,32
As mentioned above, nanostructured Ni-based materials can enhance the catalytic activity of electrochemical processes and long-term operation stability. Their physicochemical properties are determined by the synthesis conditions (reactants' concentrations, solvents used, surfactant templates, temperature, etc.) and significantly affect the material's electrical and electrochemical properties.33 Thus, Ni-based materials of various morphologies (nanowires, nanoribbons),34 nanorods, nanotubes,35 nanosheets,36 mesoporous structures,37 hollow spheres38 have been fabricated by different synthesis methods and tested for electrochemical applications.39
Herein, a hydrothermal method is used to produce Ni(OH)2 nanoplatelets. The material was characterized by scanning (SEM) and transmission electron microscopy (TEM), X-ray diffraction (XRD) analysis, as well as by Raman and X-ray photoelectron (XPS) spectroscopy. Their multifunctional electrocatalytic performance was systematically investigated for ORR, OER, and HER in alkaline media.
TEM of pure Ni(OH)2 was performed using a Hitachi H8100 200 kV transmission electron microscope with a thermionic emission gun (LaB6).
XRD equipped with Cu Kα radiation (λ = 0.15418 nm) was used to study the crystallographic structure of Ni(OH)2. The Scherrer equation (eqn (1)) was used to calculate the crystallite size,
(1) |
A micro-Raman spectrograph-Labram HR Evolution Raman spectrometer was used to characterize the Ni(OH)2 material by irradiating a 532 nm laser beam with a laser power Pl = 0.054 mW to avoid overheating on the spot.
XPS spectra for Ni(OH)2 were acquired in constant analyzer energy (CAE) mode (30 eV), using an Al (non-monochromatic) anode. The accelerating voltage was 15 kV. The deconvolution of the XPS spectra was done using Avantage® software. All spectra were referred to C 1s at 284.7 eV.
Cyclic voltammetry (CV) measurements were run in nitrogen (N2)-saturated 1 M KOH electrolyte solution at rates of 10–100 mV s−1. Linear sweep voltammetry (LSV) for ORR study was carried out in 1 M KOH solution saturated with high purity O2 at polarization rate of 5 mV s−1 and electrode rotation rates of 0, 400, 600, 800, 1000, 1200, 1400, 1600, 1800, 2000, 2200, and 2400 rpm. LSVs for OER and HER were carried out in 1 M KOH solution at a rate of 5 mV s−1 at temperatures of 25–75 °C. IR-compensation was done to minimize the influence of solution resistance (conductivity) on catalytic performance. Electrochemical impedance spectroscopy (EIS) was performed using Gamry interface 5000E for the electrode at open circuit potential (OCP) in the frequency range from 100 kHz to 0.01 Hz at an RMS amplitude of 10 mV. Chronoamperometric measurements were run under HER (−0.3 V vs. RHE) and OER (1.68 V vs. RHE) polarization conditions for 2 h.
Fig. 2a and b depicts the TEM images for Ni(OH)2 obtained at different magnifications where the nanoplatelets like morphology oriented in random directions can be seen. The nanoplatelets seem thin and uniformly distributed in good agreement with the SEM results. The selected area diffraction (SAED) pattern (Fig. 2c) reveals spotted rings from Bragg reflection from individual crystallites, confirming the poly-nano-crystallinity of the material.
The crystal structure and phase information in the Ni(OH)2 were obtained by XRD (Fig. 3a). The well-defined diffraction peaks at 2θ values of 19.2°, 33.02°, 38.48°, 52.07°, 58.97°, 62.6°, 70.42°, and 72.7° correspond to (001), (100), (101), (102), (110), (111), (103), and (201) planes, respectively, of hexagonal β-Ni(OH)2 (JCPDS no.: 014-0117)42 confirming the formation of high purity Ni(OH)2 nanocrystals. The most intense peak at 19.2°, which shows the formation of β-Ni(OH)2, was used to calculate the crystallite size using the Scherrer equation (eqn (1)). An average crystallite size of 31 nm was estimated, consistent with the SEM and TEM results.
Ni(OH)2 was further characterized by Raman spectroscopy in the 200–800 cm−1 range (Fig. 3b). Raman stretching modes in low wavenumber are generally associated with metal-oxides/hydroxides. The peak at 310.78 cm−1 was attributed to the E-type vibration of the Ni–OH lattice. The peak at 444.98 cm−1 was assigned to the Ni–O stretching (υNiO).43 In conclusion, both XRD and Raman spectrum demonstrate the formation of pure β-Ni(OH)2.
XPS analyses were carried out to further check the chemical composition, and the results are depicted in Fig. 3c and d. The Ni 2p XPS photoionizations (Fig. 3c) evidenced two major peaks with binding energies at 856.21 and 873.81 eV, corresponding to Ni 2p3/2 and Ni 2p1/2, respectively.44,45 These two peaks are separated by spin energy of 17.6 eV, which is characteristic of the Ni(OH)2 phase, in good agreement with the reported data.46 The peaks at 862.0 and 880.1 eV are associated with satellite peaks expected for Ni 2p3/2 and Ni 2p1/2 signals in the Ni 2p region. The O 1s spectra, Fig. 3d, evidenced a peak at a binding energy of 529.4 eV, which can be ascribed to the typical metal–oxygen bond of Ni–O.47 The peak at 530.5 eV is usually associated with oxygen in OH− anions.48 All the results obtained from XRD, Raman, and XPS confirm the formation of Ni(OH)2.
The accessibility of the active sites and their oxidation state will further influence the electrocatalytic reactivity. Namely, accessibility plays an essential role during OER as OER typically proceeds within a thin layer (e.g., 10 nm) of the electrocatalyst.50 Active sites in their higher oxidation state are further beneficial for OER, with this conversion to a higher oxidation state occurring at potentials lower than OER. Combining a large contact area with electrolyte and the short ion-diffusion distance, along with a low charge transfer resistance, leads to higher electrocatalytic activity. In this case, a very low charge transfer resistance value of ca. 0.7 Ω cm2 was obtained from the small semi-circle in the high-frequency region (inset of Fig. 4e), evidencing the good conductivity of the material.
LSV curve under anodic polarization conditions (Fig. 4c) reveals low OER onset potential (potential to reach a current density of 1 mA cm−2) of 1.627 V at 25 °C, further leading to an overpotential of 0.498 V to reach a current density of 10 mA cm−2. Thus, the overpotential to reach 1 mA cm−2 (reflecting the onset potential) at the herein prepared Ni(OH)2 is comparable/lower than those previously reported for Ni(OH)2 and NiM(OH)2 (M = Cr, Mn, Cu, Co, Fe, Zn) electrocatalysts ranging from 0.290 V to over 0.500 V.51 Overpotentials at 10 mA cm−2 for OER at different Ni(OH)2-based electrocatalysts have been reported to be 0.331 V for β-Ni(OH)2 NPs,16 over 0.444 V for β-Ni(OH)2 hexagonal NPs,52 0.474 V for β-Ni(OH)2 nanoplate/MWCNT,35 to 0.700 V for γ-Ni(OH)2.53 Furthermore, the reaction kinetics was found to notably improve with the increase of temperature with consequent reduction of overpotential to reach 10 mA cm−2 for ca. 80 mV.
In addition, Tafel analysis was used to evaluate the OER kinetics at this catalyst. OER on Ni(OH)2 electrocatalysts has been suggested to involve a chemical association step between OHads and OH− (eqn (3)) along with two electrochemical steps, i.e., OH− discharge (eqn (2)) and Oads− discharge (eqn (4)), followed by Oads desorption to form molecular O2 (eqn (5)):54
OH− ↔ OHads + e− | (2) |
OHads + OH− ↔ Oads− + H2O | (3) |
Oads− ↔ Oads + e− | (4) |
2Oads ↔ O2 | (5) |
4OH− ↔ O2 + 2H2O + 4e− | (6) |
Tafel slope b of 149 mV dec−1 at 25 °C decreasing to 99 mV dec−1 at 75 °C was determined for OER at Ni(OH)2, Table 1. Tafel slope values at different α-Ni(OH)2 and β-Ni(OH)2 electrocatalysts were reported to range from 42 mV dec−1 for α-Ni(OH)2 nanospheres52 and α-Ni(OH)2 hollow nanospheres33 to 161 mV dec−1 for α-Ni(OH)2 at Ni foam,55 suggesting good electrocatalytic activity of the herein prepared β-Ni(OH)2 nanoplatelets towards OER. Furthermore, a high exchange current density of ca. 85 mA cm−2 was recorded, Table 1.
T/°C | Eonset/V | η10/V | b/mV dec−1 | j0/mA cm−2 |
---|---|---|---|---|
25 | 1.623 | 0.498 | 149 | 84.6 |
35 | 1.630 | 0.486 | 131 | 85.6 |
45 | 1.625 | 0.417 | 126 | 83.0 |
55 | 1.625 | 0.453 | 119 | 81.7 |
65 | 1.626 | 0.453 | 111 | 82.4 |
75 | 1.631 | 0.453 | 99 | 83.7 |
Next, the electrocatalytic activity of Ni(OH)2 toward HER in alkaline media was evaluated. Cathodic polarization curves at 25 °C (Fig. 4d) reveal a low HER onset potential of −0.101 V so that Ni(OH)2 could reach −10 mA cm−2 at an overpotential of −0.333 V. A Tafel slope of −230 mV dec−1 was evaluated at 25 °C, decreasing to a value of −184 mV dec−1 with the increase of temperature, Table 2. Lower Tafel values have been reported upon doping Ni(OH)2 with Fe.19 Enhancement of the Ni-based materials electrocatalytic activity towards OER by introducing a second metal has been previously reported,56 resulting from the modification of intermediate bonds and electronic structures, and consequent enhancement of the electric conductivity. This suggests two possible pathways for further improvement of the herein prepared Ni(OH)2 electrocatalytic performance – doping with other metals, such as Fe or Co, or grafting onto conductive support such as reduced graphene oxide.57
T/°C | Eonset/V | η10/V | b/mV dec−1 | j0/mA cm−2 |
---|---|---|---|---|
25 | −0.102 | −0.333 | −230 | 0.402 |
35 | −0.053 | −0.256 | −211 | 0.991 |
45 | −0.028 | −0.217 | −202 | 1.051 |
55 | 0.002 | −0.181 | −203 | 1.510 |
65 | 0.047 | −0.109 | −184 | 2.913 |
75 | 0.092 | −0.065 | −204 | 4.852 |
Along with high electrocatalytic activity, stability and durability are crucial features for electrocatalyst's application in water electrolyzers. Chronoamperometric measurements in alkaline media demonstrated that β-Ni(OH)2 electrocatalyst retains its catalytic activity towards OER and HER with time (Fig. 4f). Moreover, data shows an increase in current density with time. The improvement of reaction kinetics during the long stability tests has been previously reported for OER and HER in both alkaline58,59 and acidic60 media. The increase in current density suggests that as Ni(OH)2 is preserved in alkaline media, inner regions of the material become accessible to the electrolyte with time.59
Finally, the performance of β-Ni(OH)2 for ORR was evaluated. The onset potential was found to be 0.73 V, comparable to that of freeze-dried Ni/NiOx (0.75 V) and more positive than those of freeze-dried Ni(OH)2 (0.66 V) and NiO (Fig. 5a, and Table 3).61 Similarly, ORR half-wave potential at Ni(OH)2 was found to be comparable/more positive than those of freeze-dried Ni/NiOx (0.62 V), freeze-dried Ni(OH)2 (0.56 V), and NiO (0.55 V).
Material | Morpho-logy | Crystallite/particle size/nm | Electrolyte | b/mV dec−1 | n | E1/2/V | Eon/V | Source |
---|---|---|---|---|---|---|---|---|
Ni(OH)2 | Nanoplates | 30–100 | 0.1 M KOH | 78 | 4.01 | 0.63 | 0.72 | This work |
Freeze-dried Ni(OH)2 | Layered | – | 1 M KOH | – | – | 0.56 | 0.66 | 61 |
NiO | – | – | 1 M KOH | – | – | 0.55 | 0.65 | 61 |
Freeze-dried Ni/NiOx | – | – | 1 M KOH | – | 3.75 | 0.62 | 0.75 | 61 |
α-Ni(OH)2 | Rhombohedral particle | 250–300 | 0.05–0.5 M NaOH | – | 2.9 ± 0.2 | – | – | 32 |
α-Ni(OH)2/GO | Rhombohedral particles and platelets | 150–200 | 0.05–0.5 M NaOH | – | 3.4 to 3.9 | – | – | 32 |
Ni(OH)2 | Rhombohedral particle | 250–300 | 0.5 M NaOH | — | 2.1 | — | 0.75 | 31 |
Ni(OH)2/GO | Rhombohedral particles and platelets | 150–200 | 0.5 M NaOH | – | 3.5 | – | 0.83 | 31 |
Mn2O3–NiO | – | – | 0.1 M KOH | 75, 103 | 3.89 | – | – | 62 |
NiCoO2 | – | 13 | 95 | 3.86 | 0.79 | 0.85 | 63 | |
Pt/C (40 wt% Pt) | Nanopatricles | – | 0.1 M KOH | 79, 60 | 3.96 | 0.92 | 0.96 | 62 |
Tafel analysis of the mass transfer-corrected ORR polarization curves at a rotation rate of 1800 rpm (Fig. 5a) revealed a Tafel slope of −78 mV dec−1, comparable to that of commercial Pt/C (40 wt% Pt) (−79 and −60 mV dec−1, depending on the potential region) (Table 3).62 Mn2O3–NiO also showed two Tafel regions with slopes of −75 mV dec−1 and −103 mV dec−1.62 A higher Tafel slope value of −95 mV dec−1 was reported for ORR at NiCoO2,63 which indicates the better catalytic performance of herein prepared Ni(OH)2 (Table 3).
The number of exchanged electrons, n, was determined by Koutecky–Levich analysis, and the obtained n values ranged between 3 and 4, depending on the potential (Fig. 5 and Table 1). These values suggest that ORR at Ni(OH)2 nanoplatelets proceeds by both 4e−-mechanism (eqn (7)) and 2e− mechanism (eqn (8)), with the desirable direct 4e−-reduction of O2 being the dominant one.
O2 + 2H2O + 4e− → 4OH− | (7) |
O2 + H2O + 2e− → HO2− + OH− | (8) |
Study of ORR on α-Ni(OH)2 and α-Ni(OH)2 supported on graphene oxide revealed transfer of a lower number of 2.9 ± 0.2 electrons for unsupported α-Ni(OH)2, regardless of the NaOH concentration, and transfer of 3.4 to 3.9 electrons at GO-supported α-Ni(OH)2, depending on the NaOH concentration.32 Similarly, only 2.1 electrons were exchanged during ORR at a microwave synthesized α-Ni(OH)2 and 3.5 electrons on such prepared Ni(OH)2 on graphene oxide.31 Furthermore, ORR limiting current densities at the herein studied Ni(OH)2 were observed to be higher than those recorded at microwave synthesized α-Ni(OH)2.31 Higher ORR kinetics at the GO-supported Ni(OH)2 was attributed to a lower activation energy for O2 adsorption.32 ORR was found to be a first-order reaction with respect to oxygen, i.e., the adsorption of oxygen was indicated to be the rate-determining step at both unsupported and GO-supported Ni(OH)2.32 It can be speculated that herein synthesized Ni(OH)2 nanoplatelets show higher performance than unsupported Ni(OH)2 due to a similar reaction mechanism but lower activation energy for O2 adsorption.
Considering evaluated ORR parameters, herein prepared Ni(OH)2 exhibits noteworthy ORR activity that overpasses the activity of a few previously explored Ni(OH)2 electrocatalysts in terms of higher limiting current density and higher number of exchanged electrons.31,32 Given the good electrocatalytic performance of Ni(OH)2 towards OER, HER, and ORR in alkaline media, as well as high stability, the next step would be to evaluate this material in an alkaline water electrolyzer with Ni(OH)2 as both anode and cathode electrocatalyst, and a fuel cell or metal–air battery with Ni(OH)2 cathode electrocatalyst.
Footnote |
† Equally contributed. |
This journal is © The Royal Society of Chemistry 2022 |