Previous Article in Journal
Sustainable MgO Nanocatalyst Additives for Boosting Performance and Mitigating Emissions of Used Cooking Oil Biodiesel–Diesel Blends in Compression Ignition Engines
Previous Article in Special Issue
Methanol-Tolerant Pd-Co Alloy Nanoparticles on Reduced Graphene Oxide as Cathode Catalyst for Oxygen Reduction in Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction

1
Institute of Chemistry, Technology and Metallurgy, Department of Electrochemistry, University of Belgrade, Njegoševa 12, 11000 Belgrade, Serbia
2
INS Vinca, Department of Atomic Physics, University of Belgrade, Mike Alasa 12-14, 11351 Belgrade, Serbia
3
INS Vinca, Department of Theoretical Physics and Condensed Matter Physics, University of Belgrade, Mike Alasa 12-14, 11351 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(5), 490; https://doi.org/10.3390/catal15050490 (registering DOI)
Submission received: 23 April 2025 / Revised: 14 May 2025 / Accepted: 16 May 2025 / Published: 18 May 2025
(This article belongs to the Special Issue Insight into Electrocatalysts for Oxygen Reduction Reaction)

Abstract

:
The cobalt crystalline islands (Cocryst) were electrochemically deposited onto a glassy carbon (GC) support and then modified by a facile spontaneous deposition of platinum. The electrocatalytic activity of the resulting Cocryst-Pt core-shell catalyst was evaluated for the oxygen reduction reaction (ORR) in an alkaline medium. The XRD characterization of the Cocryst-Pt islands revealed that the cobalt core had a hexagonal close-packed (hcp) crystalline structure, and that the platinum shell exhibited a crystalline structure with a preferential (111) orientation. SEM images showed that the average lateral size of the Cocryst islands was 1.17 μm, which increased to 1.32 μm after adding platinum. The XPS analysis indicated that the outer layer of the bulk metallic Cocryst islands was fully oxidized. During the spontaneous deposition of platinum, the outer Co(OH)2 layer was dissolved, leaving the cobalt core in a metallic state, while the platinum shell remained only partially oxidized. The high electrochemically active surface area of the Cocryst-Pt/GC electrode, along with a suitable crystalline structure of the Cocryst-Pt islands, contributes to enhancing its ORR activity by providing a greater number of surface active sites for oxygen adsorption and subsequent reduction. The ORR on the Cocryst-Pt catalyst occurs via a four-electron reaction pathway, with onset and half-wave potentials of 1.07 V and 0.87 V, respectively, which exceed those of polycrystalline platinum and a commercial benchmark Pt/C.

1. Introduction

Developing low-cost catalysts for an efficient oxygen reduction reaction (ORR) is crucial for advancing renewable energy technologies, including fuel cells, batteries, and water splitting [1,2,3]. Platinum-group metals are the most effective cathode catalysts, with platinum exhibiting the highest activity for oxygen reduction [4]. However, due to the high price of platinum, the key challenge for its commercial use is reducing the amount of pure platinum in a catalyst while improving its catalytic performance. In recent years, catalysts containing a small amount of Pt supported by inexpensive, electro-conductive, carbon-based materials have gained significant attention [5,6,7]. Nanostructured catalysts containing non-precious metals, which typically show low intrinsic activity for the oxygen reduction reaction, demonstrate enhanced performance when combined with other metals or appropriate supports [8]. Another promising category of materials for catalysts or supports includes transition metal oxides [9,10]. Among these, CoOx-based catalysts, such as Co3O4 and Co(OH)2, have been extensively researched for ORR and water splitting, either independently or when supported by carbon-based materials like graphene [11,12].
Among the various methods for preparing nanostructured CoOx-based catalysts [13], electrochemical deposition is frequently favored due to its simplicity and ease of use. The composition of the depositing solutions and the type of support are important factors [14]. In addition, the CoOx-species chemistry, quantity, and structure can vary depending on the electrochemical techniques employed. For potentiostatic deposition, key parameters include the applied potential and deposition time [15]. For the deposition via cyclic voltammetry, the parameters are the potential range and the number of cycles [16,17].
Various cobalt oxide (CoOx)-based catalysts have been designed and tested for their effectiveness in improving oxygen reduction reaction performance [18,19,20,21,22]. The atomic tuning of single-crystal CoO nanorods involved engineering their surface structure to create oxygen vacancies on pyramidal nanofacets, resulting in exceptionally high catalytic activity and excellent durability [18]. The CoO hybrid with nitrogen-doped carbon nanotubes exhibited a higher ORR activity compared to the Co3O4/graphene hybrid, demonstrating significant stability in a highly aggressive corrosive environment (10 M NaOH at 80 °C) [19]. Additionally, a nitrogen-doped 3D crumpled graphene-CoO hybrid displayed outstanding catalytic activity comparable to that of the benchmark Pt/C catalyst and durability that surpassed Pt/C in an alkaline medium [20]. Similar excellent ORR performance was also observed in ultra-dispersed cobalt loaded on nitrogen-doped carbon (Co-NC900) [21] and CoO@Co nitrogen-doped carbon (CoO@Co/N–C) catalysts [22]. Numerous studies indicate that combining Pt with the early transition metals (Fe, Ni, Co) can reduce costs while enhancing catalytic activity for oxygen reduction [23,24,25,26,27,28,29]. This combination leads to the formation of bimetallic catalysts, which utilize synergy and strain effects to optimize the adsorption of reactants and increase the availability of active sites on the catalyst’s surface [30].
Among the various non-precious metals, Co (Co(OH)2 and Co3O4), as a low-cost catalyst or support, combined with precious metals, is a promising material for ORR electrocatalysts. The ORR activity and durability of such Co-based catalysts can be improved by tuning their morphology and combining them with Pt nanoparticles [28,29,30,31]. Oxygen reduction was studied on carbon-supported Pt–Co nanoparticles with low Pt loading in an alkaline solution. Despite the low platinum content, significant activity was observed in the Pt-rich shell and Pt-Co core particles obtained through potential cycling [32]. Various platinum-cobalt alloy nanoparticles, prepared using the nanocapsule method, were found to exhibit higher ORR catalytic activity and stability in alkaline electrolytes than in acidic [33]. Bimetallic platinum-cobalt nanoparticles supported on reduced graphene oxide, Pt3Co/rGO, exhibited superior oxygen reduction activity compared to other tested catalysts, including Pt/C, Pt/rGO, and Pt3Ni/rGO, in alkaline solutions [34]. In a separate study, PtCo alloyed nanoparticles supported on thiolated graphene (referred to as ER/PtCo-tG) were electrochemically reduced to improve their catalytic activity. The enhanced ORR activity can be attributed to the well-dispersed PtCo nanoparticles, the effective formation of the PtCo alloy, and the presence of Co(OH)2 [35].
In this work, we investigated glassy carbon (GC)-supported Co-Pt catalysts for oxygen reduction in an alkaline medium. The electrochemically deposited cobalt crystalline catalysts, obtained by applying various constant potentials and durations, were explored as substrates for further modification by a spontaneously deposited platinum layer. Although the spontaneous deposition of Pt on bare GC is nearly hindered [36], it occurs with remarkable ease on GC covered with previously electrodeposited Co crystalline islands. The layer of platinum shells the cobalt crystalline islands, thus preserving their structural integrity and enhancing their catalytic performance during oxygen reduction. The most ORR active GC-supported Cocryst and Cocryst-Pt catalysts were characterized using cyclic voltammetry (CV), X-ray diffraction (XRD), scanning electron microscopy (SEM), and X-ray photoelectron spectroscopy (XPS). A rotating disc electrode (RDE) arrangement was explored for investigations of the oxygen reduction reaction on Cocryst/GC and Pt-Cocryst-Pt/GC electrodes using electroanalytical techniques, including linear sweep voltammetry (LSV) and chronoamperometry (CA).

2. Results and Discussion

2.1. The Preparation and Characterization of Co/GC and Co-Pt/GC Electrodes

2.1.1. The Preparation of Co/GC and Co-Pt/GC Electrodes

The electrochemical deposition of Co onto a GC substrate was carried out from a deaerated 10−4 M CoSO4 aqueous solution at various deposition times and different constant potentials, as described in the Supplementary Materials and illustrated in Figure S1. The cyclic voltammetry characterization of various Co/GC electrodes, conducted in a deaerated 0.1 M NaOH, along with preliminary testing of their ORR activity in an oxygen-saturated 0.1 M NaOH, is detailed in the SI and depicted in Figure S2. Since the Co/GC electrode produced by Co deposition at a constant applied potential of −0.36 V for 30 min demonstrated the best ORR activity, it will be further utilized as a substrate for subsequent spontaneous Pt deposition. The chronoamperometry curve (Figure S3) and the calculation of the Co metal loading achieved during the 30 min deposition (0.3 mg cm−2) are presented in the Supplementary Materials.
According to the Nernst equation and the standard potential for Co2+/Co of −0.28 V vs. SHE [37], the calculated equilibrium potential for Co electrodeposition from a 10−4 M Co2+ solution at a pH of 6.1 is − 0.34 V vs. RHE. A slightly lower selected potential of −0.36 V (−0.71 V vs. SHE), according to the Pourbaix diagram [38,39], indicates that only crystalline metallic Co is deposited, according to Equation (1). This electrode obtained after 30 min Co electrochemical deposition at −0.36 V will be referred to as Cocryst/GC or the Cocryst catalyst.
C o 2 + + 2 e C o c r y s t 0
In addition, in contact with a close to neutral Co depositing solution, the outer layer of the deposited cobalt islands undergoes oxidation to Co(OH)2, with a simultaneous hydrogen evolution. The Cocryst islands will be referred to as Cocore_Co0, or Cocore_Co(OH)2, indicating that the Co atoms in the outer layer are in a zero-oxidation state or oxidized.
C o c o r e _ C o 0 C o c o r e _ C o O H 2 + H 2
Co-Pt catalysts were prepared by immersing the most ORR-active Cocryst/GC electrode in a deposition solution consisting of 1 mM H2PtCl6 and 0.05 M H2SO4 for various times. This is described in the SI and illustrated in the CVs shown in Figure S4a. According to their corresponding ORR polarization curves in Figure S4b, the best ORR activity is achieved for 60 min Pt deposition. This electrode will be referred to as the Cocryst-Pt/GC or as the Cocryst-Pt catalyst.

2.1.2. The Electrochemical Characterization of Cocryst/GC and Cocryst-Pt/GC Electrodes

The CV of Cocryst/GC obtained for Co deposition at −0.36 V (Figure 1a) exhibits several distinct properties during the anodic scan characteristic of the bulk cobalt electrode. These include an extended peak at approximately 0.42 V with a shoulder at a lower potential attributed to the stripping of Co(OH)2 [3] along with several anodic peaks at higher potentials associated with the transition to higher oxidation states of cobalt, including Co(II)/Co(III), and the formation of Co3O4 and CoOOH [39,40,41,42]. Additionally, there may also be peaks related to the Co(III)/Co(IV) transition [43].
The first anodic peak (from 0.0 to 0.9 V), the stripping of the Co(OH)2 outer layer, may be used to determine the electrochemically active surface area (ECSA) of the Cocryst/GC electrode. The charge obtained by integrating the Co/Co(OH)2 peak is 17.8 mC cm−2. By dividing this value with the calculated charge passed during the stripping of the first Co(OH)2 monolayer of 406 μC/cm2 (considering the exchange of 2e per single cobalt atom, a Co(OH)2 unit cell dimension of 0.32 x 0.32 nm, and the geometric surface area of 0.196 cm2), the resulting ECSA value for Cocryst/GC is 8.6 cm2.
Figure 1b shows changes in the open circuit potential (OCP) during Pt spontaneous deposition. The starting OCP value for the Cocryst/GC electrode is −0.02 V vs. RHE (−0.08 V vs. SHE), confirming that at this potential, according to the Pourbaix diagram [38,39], the surface of Cocryst islands is covered with Co(OH)2. Such islands consisting of a metallic Co core encapsulated by a Co(OH)2 shell will be referred to as C o c o r e 0 _ C o ( O H ) 2 , s h e l l . Analogous core-shell Co_Co(OH)2 nanoparticles were reported in ref. [39]. This Co(OH)2 shell undergoes a dissolution in an acid Pt depositing solution (pH 1), leaving the metallic Co0 place free to become simultaneously occupied by the adsorbed P t C l 6 , a q 2 anions from the depositing solution according to Equations (3) and (4):
C o c o r e 0 _ C o ( O H ) 2 , s h e l l + 2 H + C o c o r e 0 + C o a q 2 + + 2 H 2 O
C o c o r e 0 + P t C l 6 , a q 2 C o c o r e 0 _ P t C l 6 , a d s 2
Due to the higher Nernst potential for Pt4+/Pt0 (0.07 V vs. SHE, see ref. [36]) than for Co2+/Co0, the Pt4+ from the adsorbed P t C l 6 , a d s 2 anion is spontaneously reduced to metallic Pt, while the two Co0 atoms from the surface of a Co core are simultaneously oxidized to Co2+, thus forming the islands consisting of metallic Co-core and metallic Pt-shell:
C o c o r e 0 _ 2 C o 0 _ P t C l 6 , a d s 2 C o c o r e 0 _ P t s h e l l 0 + 2 C o a q 2 + + 6 C l a q
A slight increase in the OCP over time indicates that the surface Co(OH)2 oxide is gradually dissolving and subsequently being replaced by spontaneously deposited metallic Pt. Upon completing this process, there is a further sharp increase in the OCP to 0.92 V, corresponding to the OCP of bare polycrystalline Pt in 0.05 M H2SO4 (Figure S5). According to the CV of bare Pt(poly) in 0.05 M H2SO4 (insert in Figure S5), at the potential corresponding to this OCP value, Pt is in the oxidized state. With prolonged spontaneous Pt deposition, the OCP remains constant, indicating that Pt is additionally deposited on the existing Pt shell, resulting in an increased thickness of the Pt layer.
Figure 1c,d display the CVs of Cocryst-Pt/GC, Ptspont/GC, and Pt(poly), respectively, in a 0.1 M NaOH solution, recorded after holding the potential for 5 min at 0.2 V to reduce Pt oxide to its metallic state, and after the first scan in wider potential limits to remove the traces of chlorides and other impurities. The CVs of Cocryst-Pt/GC and Pt(poly) exhibit all the typical features of polycrystalline Pt in 0.1 M NaOH, consistent with previous reports [44]. The hydrogen adsorption/desorption process occurs within the potential range of 0.05 V to 0.18 V, featuring two pairs of peaks in both the cathodic and anodic directions. Oxidation begins at approximately 0.60 V with the formation of Pt(OH)2 and PtO at higher potentials, showing a distinct peak at 0.86 V. The reduction peak is observed at 0.70 V. In contrast, the CV of the Ptspont/GC electrode, obtained by a spontaneous Pt deposition for 60 min (Figure 1d), indicates that there is little to no Pt on the GC support, which is obvious when compared to the current densities in the CV of bare Pt. In addition, much higher current densities for Cocryst-Pt/GC compared to bare Pt(poly) suggest much higher electrochemically active surface area (ECSA). The charge obtained by integrating hydrogen desorption peaks is 1.53 mC cm−2 for Cocryst-Pt/GC, and 259 μC cm−2 for the Pt(poly) electrode used in the work. Considering the commonly accepted value for a hydrogen monolayer on a polycrystalline Pt of 210 μC/cm2, the ECSAs for the Cocryst-Pt/GC and Pt(poly) are 1.43 cm2 and 0.24 cm2, respectively.

2.1.3. Structural, Morphological, and Chemical Characterization of the Cocryst/GC and Cocryst-Pt/GC

XRD, SEM, and XPS measurements were performed on the as-prepared electrodes. After the deposition process is finished, the electrodes are transferred through the air before additional measurements. Since both Cocryst/GC and Cocryst-Pt/GC electrodes contain oxides formed in contact with the solution, their outer spheres are also susceptible to further oxidation in air.
Figure 2 shows XRD profiles and corresponding SEM images of electrochemically deposited Cocryst/GC and Cocryst-Pt/GC, acquired through additional spontaneous Pt deposition. Amorphous peaks of glassy carbon dominate both XRD profiles of Cocryst/GC (Figure 2a) and Cocryst-Pt/GC (Figure 2b). For the Cocryst/GC, the diffraction pattern displays three peaks centered at 41.70 (100), 44.50 (002), and 47.50 (101), the signature peaks for the metallic hexagonal close-packed (hcp) Co phase (PDF 1-1256; hexagonal s.g. P63/mmc no.194) [45,46]. There are no traces of face-centered cubic (fcc) (signature peak at 520) structure [46] or the Co oxide (that would be at 36.3 (111) and 42.3 (200)) [47]. This result indicates that the electrochemically deposited Cocryst islands are composed of elemental crystalline Co within their volume, with the exposed preferential signature (100) and (002) peaks. This is consistent with the observation for Co nanoparticles electrodeposited at the same applied potential of −0.36 V [48]. Due to the size of the deposited Co crystallites, the exposure of the (110) orientation is likely associated with the preferential alignment of their edges.
For the Cocryst-Pt/GC, in addition to the peaks of glassy carbon, small but noticeable peaks of metal Pt (PDF 1-1194; cubic s. g. Fm-3m no. 225) appear in the XRD profile, with no traces of Co metal or Co oxide phases, suggesting a very thick Pt shell. The preferred orientation of the Pt shell includes (111) and (110), consistent with the face-centered cubic (fcc) crystal structure of platinum. Moreover, the hcp Co surface of (100) orientation is structurally similar to the fcc-Pt(111), both exhibiting a hexagonal structure of surface atoms, suggesting that the Pt layer is expected to exhibit a preferential (111) orientation.
Figure 2c,d show SEM images of the Cocryst/GC (left column) and Pt-Cocryst-Pt/GC surface (right column), showing surface topography, revealing that in both cases, the deposited larger Cocryst and Cocryst-Pt hexagonally shaped crystalline islands are randomly distributed over the GC surface. In addition, the image of Cocryst/GC reveals a layer of the deposited cobalt species between the Cocryst islands. In contrast, the image of Pt-Cocryst-Pt/GC reveals only bare GC support between the deposited Cocryst-Pt islands, meaning that the previously deposited CoOx layer was dissolved during spontaneous Pt deposition from an acidic deposition solution and that Pt is deposited on the remaining bare metallic Cocryst islands. A spontaneous deposition of Pt on bare GC support can be regarded as negligible. The average lateral sizes of individual Cocryst and Cocryst-Pt islands are 1.17 μm and 1.32 μm, respectively, as shown in the island size distribution diagrams in Figure 2e,f, obtained from the images with a lower magnification as illustrated in the SI, Figure S6.
The chemical composition analysis was performed using energy dispersive X-ray spectroscopy (EDS). This hcp structure of the deposited Co crystalline islands aligns well with the hexagonal shape of the Cocryst islands in the corresponding SEM image (Figure 3a). The SEM image in Figure 3b reveals that the hexagonal shape of the deposited islands is preserved upon the addition of a Pt shell. EDS analysis (Figure 3c,d) confirms that Cocryst/GC consists of carbon, oxygen, and cobalt, while Cocryst-Pt/GC consists of carbon, oxygen, cobalt, and platinum.
The XPS analyses presented in Figure 4 provide information on the surface composition of the outer layer of Cocryst islands and of the Pt shell of the Cocryst-Pt core-shell islands. The XPS survey spectrum of Cocryst/GC, Figure 4a, provides a broad overview of the electrode surface composition, revealing multiple peaks corresponding to the elements present, including Co, C, and O. No other elements are detected in the spectrum.
The Co 2p spectrum for Cocryst/GC, Figure 4b, displays prominent peaks at binding energies of 781.6 eV and 797.4 eV, corresponding to the Co 2p3/2 and Co 2p1/2 levels, respectively. These peaks are ascribed to the cobalt oxide species present in the Co2+ oxidation state [49,50,51,52], most likely Co(OH)2 [49,50], or as a mixture of CoO and Co(OH)2 [50]. Additionally, two satellite peaks are observed at 786.3 eV and 803.6 eV, characteristic of Co2+ species resulting from multiplet splitting, which arises from the interaction of the 2p core hole with the unfilled 3d states of the Co2+ ion. These satellites are commonly observed in cobalt oxides, further confirming the presence of Co2+ in the Co(OH)2 phase [49]. No Co3+ [51,52] and metal Co0 [49] are observed in the spectrum.
The O 1s spectrum for Cocryst/GC (Figure 4c) shows two distinct peaks. The peak at 530.3 eV is attributed to the oxygen in Co(OH)2-CoO (7.4% of O 1s emission), confirming the presence of cobalt oxide on the surface. The second peak, at 532.0 eV, can be attributed to adsorbed oxygen species (O, O2−, O2), accounting for a large amount of oxygen detected in the survey spectrum (92.6% of O 1s emission). In addition, based on a small C 1s signal in the survey spectrum, the same peak at 532.0 eV can be attributed to C-O bonds from the GC support, which are located at a similar binding energy [51,52], although in a much smaller amount.
The XPS survey spectrum of Cocryst-Pt/GC, Figure 4d, displays multiple peaks corresponding to Pt, C, O, and Cl. In this case, Co cannot be detected, as most Co-crystalline islands are encapsulated by platinum, which is preferentially deposited on the pre-existing cobalt crystals. Since XPS is a surface-sensitive technique, it cannot detect cobalt beneath the platinum layer. The peak corresponding to Cl is attributed to the chlorine from the precursor H2PtCl6 chemical used for platinum deposition.
The Pt 4f spectrum of Cocryst-Pt/GC, Figure 4e, exhibits six peaks, corresponding to three distinct components in the Pt 4f7/2 and Pt 4f5/2 regions. These peaks are assigned to distinct chemical states of platinum present on the surface. The peaks at 75.2 eV and 71.8 eV (47.7% of Pt 4f emission) are associated with Pt0, indicating the presence of platinum in its metallic state [36,53]. The peaks at 75.7 eV and 72.8 eV (48.6% of Pt 4f emission) correspond to PtOx, which represents platinum in an oxidized state, or platinum species that have undergone partial oxidation during the exposure of the electrode in air, possibly as platinum oxide, platinum hydroxide Pt(OH)2 [36,53]. A roughly equal amount of platinum is in the metallic and oxidized states. The remaining small peaks at 78.3 eV and 73.8 eV (3.7% of Pt 4f emission) are attributed to PtCl4, indicating platinum in a chloro-complex form, likely due to residual chlorine from the H2PtCl6 precursor chemical used during platinum deposition [36,53].
The spectrum of O 1s for Cocryst-Pt/GC, Figure 4f, shows two distinct peaks at 531.7 eV and 532.6 eV, corresponding to oxygen bound to Pt in various PtOx states (67.4% of O 1s emission) in either various adsorbed oxygen species (O, O2−, O2−) or C-O bonds from the GC substrate (32.6% of O 1s emission), respectively [36,53]. Compared to the O 1s spectrum of Cocryst/GC, the Cocryst-Pt/GC spectrum shows significantly less oxygen overall, with the C 1s peak exhibiting much higher intensity. This suggests that the O 1s signal in the Cocryst-Pt/GC originates mainly from the GC support rather than from adsorbed oxygen species. The increase in the relative intensity of the C 1s peak indicates that the carbon support dominates the surface more. In contrast to the previous case, where the GC surface between Co crystalline islands was also covered by the deposited Co oxide, during spontaneous Pt deposition, these Co oxides are dissolved in acidic depositing solution, leaving partially exposed bare GC substrate among the Cocryst-Pt islands. This increases the amount of oxygen bound to the carbon from the GC support and reduces the amount of oxygen bound to surface metal oxide species, specifically PtOx.
The atomic and weight percentages of the constitutive elements for Cocryst/GC and Cocryst-Pt/GC are shown in Table 1, as estimated from the relative intensity of their corresponding peaks in the survey spectra. It is important to note that in the case of Cocryst-Pt/GC, cobalt is not detected by XPS, and its amount cannot be estimated. This indicates that Cocryst islands are fully covered with a Pt layer. According to the island size distribution (see Figure 2e,f), the average sizes of the Cocryst and Cocryst-Pt islands are 1.17 and 1.32 μm, respectively. The difference of 150 nm suggests that the average thickness of the Pt layer is 75 nm. Since XPS enables analysis of samples to a depth of 3–10 nm, the Co core lies beyond this range, allowing only the outer Pt layer to be detected and analyzed.

2.2. Oxygen Reduction Reaction on Cocryst/GC and Cocryst-Pt Catalysts

The polarization curves for the oxygen reduction reaction, recorded in an oxygen-saturated 0.1 M NaOH solution and the corresponding Koutecky–Levich (K–L) plots, are shown in Figure 5.
The ORR curves in Figure 5a,b were recorded in the cathodic scan direction for five rotation rates. Generally, the oxygen reduction on CocrystPt/GC occurs at more positive potentials and with higher current densities than on Ccryst/GC. The onset (Eonset) and half-wave (E1/2) potentials for ORR on Ccryst/GC are 0.82 and 0.62 V, respectively. On the other hand, for ORR on CocrystPt/GC, Eonset and E1/2 are 1.07 and 0.87 V, respectively.
Koutecky–Levich plots, Figure 5b,d, are constructed using data from polarization curves (Figure 5a,c) when the inverse current density (1/j) is plotted as a function of the inverse of the square root of the rotation rate (ω1/2).
For Cocryst/GC, Figure 5b illustrates that K–L diagrams are parallel across the entire ORR potential region, indicating that the reaction adheres to first-order kinetics with respect to molecular oxygen. The calculated number of exchanged electrons remains consistently four, irrespective of variations in potential. These results demonstrate that the Cocryst/GC electrode facilitates ORR through a 4e reaction pathway. Since in the entire ORR potential region, the surface is covered with Co(OH)2, it may be assumed that the OH acts as a catalyst for the adsorption of oxygen and its subsequent reduction, like in the case of Au(100) single crystal [54].
The ECSA provided above for Cocryst/GC, estimated from CV in Figure 1a and corresponding to the stripping of the first Co(OH)2 layer, reflects the state of the as-prepared surface. However, at the ORR onset potential of 0.82 V, the surface has already transformed from its initial state (as illustrated in Figure S7 by the repeated cycling in the potential region from 0.0 to 0.9 V). This indicates that the ECSA corresponding to the onset of ORR differs from the initial one at around 0.0 V, which is demonstrated by the difference in double-layer capacity (Figure S8a,b). Considering that the Cdl value for bare Co is 40 μF cm−2 [55], the obtained ECSA for the potential region from −0.02 to 0.08 V (0.03 V midpoint potential) is 9.8 cm2, while for the potential region around ORR onset potential from 0.79 to 0.85 V, it is 2.9 cm2 (see Figure S9a,b). This implies that, at more positive potentials, the surface is partially passivated due to the formation of higher oxides, resulting in a lower availability of Co(OH)2 surface sites, which act as catalysts for oxygen adsorption and its subsequent reduction.
For Cocryst-Pt/GC, Figure 5d, the linear K–L diagrams suggest first-order kinetics of the ORR concerning molecular oxygen. Unlike in the previous case, slight non-parallelism of the K–L plots at higher potentials indicates a change in the total number of electrons exchanged with potential. In these activation and mixed control regions, the ORR occurs on the oxidized Pt layer, suggesting that the presence of PtOH on the surface partially hinders oxygen adsorption [56]. In contrast, in the diffusion control region at lower potentials, the K–L curves become parallel, indicating that the total number of electrons exchanged remains constant and does not vary with potential. The total number of electrons exchanged is four, and the ORR follows a 4e reaction pathway.
Like in the previous case, considering that the Cdl value for bare Pt is 60 μF cm−2 [55], the ECSAs for Cocryst-Pt/GC obtained from CV in Figure 1c and from Cdl around the ORR onset potential (from 1.02 to 1.2 V, midpoint potential 1.07 V, see Figure S9c,d) differ. The ECSA value of 5.2 cm2 obtained from Cdl is significantly higher than the 1.43 cm2 from the hydrogen desorption peak, contributing to the enhanced ORR activity.

Comparison of the ORR Catalytic Activity and Stability of Cocryst-Pt/GC with That of Ptspont/GC, Pt(poly), and Cocryst/GC

Figure 6 shows LSV curves for oxygen reduction in 0.1 M NaOH recorded in the cathodic direction at a rotation rate of 1600 rpm, along with the corresponding Tafel plots for the Ptspont/GC, Pt(poly), Cocryst/GC, and Cocryst-Pt/GC electrodes.
The oxygen reduction polarization curves (Figure 6a) indicate that the ORR catalytic activity of Ptspont/GC is relatively low, with an onset potential (Eonset) of 0.72 V and a half-wave potential (E1/2) of 0.58 V. This low ORR activity, which is only slightly better than that of bare GC (shown in Figure S1b), is attributed to the negligible Pt content. In contrast, the Cocryst/GC electrode demonstrates much higher catalytic activity, although it still does not reach that of Pt(poly) (Eonset = 0.98 V, E1/2 = 0.84 V). On the other hand, the Cocryst-Pt/GC shows a significant enhancement in catalytic activity, surpassing that of Pt(poly).
The corresponding Tafel slopes are shown in Figure 6b. The slope of −100 mV dec−1 for the ORR on Ptspont/GC over the entire potential range matches the value obtained for bare GC [57], as well as for Cocryst/GC. The slope of −100 mV dec−1, obtained for the ORR on Cocryst/GC, is comparable to that of various carbon-based electrocatalysts featuring Co nanoparticles [21,51,58,59]. For the Pt(poly) and Pt-Cocryst/GC electrodes, two different Tafel slope values were identified at low and high overpotentials. The −60 and −120 mV dec−1 slopes for Pt(poly) align with previous reports [36]. The two slopes at high and low overpotentials reflect the changes in adsorption conditions, transitioning from a Temkin (−60 mV dec−1) to a Langmuir (−120 mV dec−1) model, as the transfer of the first electron to O2 adsorbed on the surface is the rate-determining step, as already discussed in ref. [36].
The parameters for the GC-supported Cocryst, Cocryst-Pt, and Ptspont catalysts compared to bare Pt(poly), including ECSA, Eonset, E1/2, and Tafel slope for the ORR in 0.1 M NaOH solution at 1600 rpm, are summarized in SI, Table S1.
Figure 7 illustrates the long-term stability testing of the Cocryst-Pt catalyst during the ORR in oxygen-saturated 0.1 M NaOH. The testing involved recording a chronoamperometry curve over 5 h at a constant applied potential and 2000 LSV curves during continuous cycling within the ORR potential range at a scan rate of 100 mV s−1. In all cases, the rotation rate was 1600 rpm.
The CA curves (Figure 7a) illustrate the long-term stability performance of the Cocryst-Pt core-shell islands compared to Cocryst islands and Pt(poly). At an applied potential of 0.87 V, the ORR exhibits significantly higher current density for the Cocryst-Pt than for Pt(poly). Initially, the current density decreases, but then stabilizes during prolonged measurements. However, a slight decline in activity is observed for extended durations of the ORR on Cocryst-Pt. A similar trend is noted for the ORR on Cocryst alone (recorded at the potential of 0.67 V), although there is a greater decrease in activity during the first hour.
Figure 7b,c show LSV curves for ORR on Cocryst-Pt catalyst before and after the stability measurements and before and after 2000 cycles (performed at a scan rate of 100 mV s−1), respectively, recorded at a scan rate of 50 mV s−1. In both cases, LSV curves exhibited a slight decline in stability (similar to that of bare Pt(poly), as illustrated in Figure S10a). In contrast, the LSV curves for ORR on Cocryst islands (Figure S10b) demonstrate a significant ORR activity decline, likely due to the formation of higher Co oxide, which contributes to increased surface passivation. The better stability of the Cocryst-Pt catalyst suggests the protection of the Cocryst core by a Pt layer. However, further work is necessary to improve the long-term stability of the Cocryst-Pt catalyst, primarily focusing on preventing the degradation and dissolution of the outer Pt layer.
Due to the large size of Cocryst islands at the micrometer scale and the thickness of the platinum layer of a few tenths of nanometers, no electronic effects are expected to play a role in the increased ORR activity of the Cocryst-Pt core-shell islands compared to Cocryst alone. Based on the presented results, it is reasonable to conclude that, besides the higher intrinsic activity of platinum compared to cobalt, the primary reason for the enhanced ORR activity is the favorable surface chemical composition and structure of the Cocryst islands. This Co surface chemistry and structure facilitate the spontaneous deposition of Pt. Moreover, a hexagonal structure of both Co(100) and Pt(111) surface atoms promotes the growth of a crystalline Pt(111) layer and maintains the crystalline integrity of the electrochemically deposited Cocryst islands. This leads to the significantly increased ECSA of Cocryst-Pt core-shell islands compared to both Cocryst islands alone and Pt(poly), meaning a greater number of surface active sites for oxygen adsorption and its subsequent reduction. As a result, the electrochemically active surface area of the Cocryst-Pt core-shell islands is significantly greater than that of polycrystalline platinum. Such a high ECSA corresponds to many platinum surface active sites available for oxygen adsorption and subsequent reduction, thus enhancing the overall ORR efficiency.
In addition, the results obtained in this work are compared with the literature data for ORR in both alkaline and acid media on various Co-Pt catalysts supported by carbon-based materials, as well as with the benchmark Pt/C catalyst, as summarized in Table 2.
The ORR activity of the obtained GC-supported Cocryst-Pt catalyst is among the best concerning the initial potential of 1.07 V, which is within the range of the best Pt-Co catalysts reported in refs. [34,60], surpassing the activity of Pt(poly) and Pt/C [60,70]. The Tafel slopes of −63 and −120 mV dec−1 for the Pt-Cocryst/GC are consistent with those previously reported in refs. [60,61,62], although a lower value is reported in refs. [63,64].

3. Materials and Methods

3.1. Catalyst Preparation

The electrochemical deposition of cobalt onto a glassy carbon supporting electrode was conducted in a three-electrode cell using a 10−4 M CoSO4∙7H2O aqueous solution. Different constant deposition potentials and times were applied to create Co/GC electrodes. Additionally, platinum was spontaneously deposited by immersing previously prepared Co/GC electrodes into a 10−3 M H2PtCl6 + 0.05 M H2SO4 solution for various durations.

3.2. Material Characterization

XRD measurements were performed using a Philips PW 1050 X-ray powder diffractometer (Philips, Delft, The Netherlands). The measurements utilized Ni-filtered Cu Kα radiation and employed Bragg–Brentano focusing geometry. Diffraction intensity data were recorded over the 2θ range of 10–80° with a step size of 0.02° and a counting time of 5 s per step.
SEM images and EDS spectra were recorded using a FESEM microscope (FEI Scios 2, Thermo Fisher Scientific, Waltham, MA, USA) at a pressure of 1 × 10−6 mbar and an electron beam voltage of 10 kV. The duration for spectrum acquisition from the selected micro areas was 30 min.
Survey and high-resolution XPS spectra were recorded using SPECS Systems (SPECS Surface Nano Analysis GmbH, Berlin, Germany) with XP50M X-ray source for Focus 500 and PHOIBOS 100/150, using AlKα source (1486.74 eV) at a 12.5 kV, and 32 mA at a pressure of 9 × 10−9 mbar. C 1s peak at 284.5 eV was used as a reference for all peak positions.

3.3. Electrochemical Measurements

The PINE potentiostat (PINE Instruments Co., Grove City, PA, USA) and a three-electrode electrochemical cell were employed for electrochemical measurements. Different Co and Co-Pt catalysts supported by GC were prepared and used as the working electrodes. A Pt-wire served as the counter, while an Ag/AgCl, 3 M KCl was utilized as the reference electrode. All potential values were recalculated relative to the reversible hydrogen electrode (RHE).
The electrochemical characterization of the electrodes was conducted using cyclic voltammetry in a deaerated 0.1 M NaOH solution. Their ORR electrocatalytic activity was investigated in an oxygen-saturated 0.1 M NaOH solution through linear sweep voltammetry in a rotating-disc electrode setup.
The Koutecky–Levich (K–L) plot presents the relationship between the inverse current density value (1/j) and the inverse square root of the rotation rate (ω−1/2).
The Koutecky–Levich equation was used for a detailed ORR analysis:
1/j = 1/jk + 1/jl = 1/jk + 1/Bω1/2
where j is the measured current density and jk and jl represent the kinetic and diffusion current density, respectively. B is a constant expressed as:
B = 0.62 nF C o 2 D o 2 2 / 3 ν 1 / 6
where n is the total number of exchanged electrons, F is the Faraday constant (96,485 C mol−1), C o 2 is the oxygen solubility (1.22 × 10−3 mol L−1 [36]), D o 2 is the oxygen diffusivity (1.90 × 10−5 cm2 s−1 [36]), and ν is the kinematic viscosity (0.01 cm2 s−1 [36]) of the 0.1 M NaOH solution. The value of the constant B determined experimentally from the K–L diagram was used to calculate the total number of electrons exchanged.

3.4. Chemicals

Glassy carbon discs (Alfa Aesar GmbH & Co KG, Karlsruhe, Germany), mechanically polished with alumina powder of 1, 0.3, and 0.05 µm particle sizes, were used as supporting electrodes. A glassy carbon disc with a diameter of 5 mm (geometric area of 0.196 cm2) was mounted in a Teflon holder for electrochemical measurements. Additionally, a 7 mm diameter disc (geometric area of 0.385 cm2) was used for ex situ characterization.
CoSO4∙7H2O and H2PtCl6 (both from MaTech GmbH, Jülich, Germany), H2SO4 (Merck, Germany), and ultrapure Milli-Q water were used to prepare the Co and Pt deposition solutions. NaOH pellets (from Merck, Darmstadt, Germany) and Milli-Q water were used for the working solution for electrochemical measurements. The solutions were either deaerated with nitrogen (N2, 99.9995%, Messer, Frankfurt, Germany) or saturated with oxygen (O2, 99.9995%, Messer, Frankfurt, Germany).

4. Conclusions

The results presented in this study demonstrate a straightforward method for creating a Co-Pt core-shell catalyst by electrochemically depositing cobalt onto a glassy carbon support, followed by a spontaneous deposition of platinum. The Cocryst and Cocryst-Pt catalysts that have shown the best ORR activity were characterized, and their electrochemical performances were investigated in 0.1 M NaOH in detail.
For Cocryst/GC, the XRD indicated the crystalline nature of the deposited Co islands with the preferential (100) surface structure, while SEM analyses revealed an average lateral size of 1.17 μm. XPS detected cobalt in only the Co2+ oxidation state, suggesting that Cocryst islands consisted of a metallic Co-core and Co(OH)2 shell. After adding Pt, XRD indicated that the metallic hexagonal close-packed Co core promotes the formation of a crystalline Pt shell with a preferred (111) orientation. SEM analysis revealed an increase in the average size of the Cocryst-Pt islands to 1.32 μm. XPS detected only Pt, confirming a full coverage of the Co core with a thick Pt shell.
RDE results on the ORR activity of the Cocryst-Pt/GC electrode demonstrated that the reaction proceeds through a 4e-reaction mechanism with an onset and half-wave potentials of 1.07 and 0.87 V, respectively, and a Tafel slope of 63 mV dec−1. Due to the high intrinsic activity of the platinum shell and the large electrochemically active surface area, the resulting Cocryst-Pt/GC catalyst exhibited enhanced catalytic activity for the oxygen reduction reaction, exceeding that of a polycrystalline platinum and a benchmark Pt/C.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15050490/s1, Figure S1: The electrochemical deposition of cobalt on the GC support from a deaerated 10−4 M CoSO4 aqueous solution: (a) CV curve recorded at a scan rate of 50 mV s−1; (b) CA curves recorded at different deposition potentials; Figure S2: Electrochemical performance of Co/GC electrodes: (a) CVs recorded after a 5 min deposition of Co at various potentials in deaerated 0.1 M NaOH; (b) corresponding ORR LSV curves in oxygen-saturated 0.1 M NaOH; (c) CVs of 30 min Co/GC obtained for various Co deposition times; (d) corresponding ORR LSV curves; Figure S3: CA curve recorded at −0.36 V during 30 min cobalt deposition; Figure S4: Electrochemical performance of the Pt/Co electrodes: (a) CVs in deaerated 0.1 M NaOH; (b) LSV curves in oxygen-saturated 0.1 M NaOH. Scan rate 50 mV s−1; Figure S5: OCP curve of Pt(poly) in 0.05 M H2SO4, and (insert) corresponding CV recorded at 50 mV s−1; Figure S6: SEM images of Cocryst/GC (left column) and Cocryst-Pt/GC (right column) electrodes: (a,c) magnification = 2000×; (b,d) magnification = 5000×; Figure S7: CVs of Cocryst/GC in 0.1 M NaOH, recorded during repeated cycling from 0.0 to 0.9 V. Scan rate 50 mV s−1; Figure S8: Double layer capacity determination for Cocryst/GC electrode in 0.1 M NaOH solution: (a) CVs recorded in the double layer region from −0.02 to 0.08 V with different scan rates; (b) corresponding Δj/2 vs. scan rate plot; Figure S9: Double layer capacity determination for Cocryst/GC and Cocryst-Pt/GC electrodes in 0.1 M NaOH solution: (a,b) CVs recorded in the potential region around the ORR onset potential with different scan rates; (e,f) corresponding Δj/2 vs. scan rate plots; Table S1: The activity of Cocryst-Pt, Cocryst, and Ptspont catalysts compared to bare Pt(poly) for ORR in 0.1M NaOH solution; Figure S10: LSV curves recorded before and after 5 h ORR stability measurements in 0.1 M NaOH of: (a) Pt(poly); and (b) Cocryst/GC. Scan rate 50 mV s−1. References [14,39,55] are cited in the supplementary materials.

Author Contributions

Conceptualization, J.G. and S.Š.; investigation, J.G., L.R., V.R. and M.M; writing—original draft preparation, J.G., L.R. and M.M.; writing—review and editing, J.G. and S.Š.; supervision, S.Š. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education, Science and Technological Development of the Republic of Serbia (Contract No: 451-03-136/2025-03/200026).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zaman, S.; Huang, L.; Douka, A.I.; Yang, H.; You, B.; Xia, B.Y. Oxygen reduction electrocatalysts toward practical fuel cells: Progress and perspectives. Angew. Chem. Int. Ed. Engl. 2021, 9, 17832–17852. [Google Scholar] [CrossRef]
  2. Zhao, J.; Wei, D.; Zhang, C.; Shao, Q.; Murugadoss, V.; Guo, Z.; Jiang, Q.; Yang, X. An overview of oxygen reduction electrocatalysts for rechargeable zinc-air batteries enabled by carbon and carbon composites. Eng. Sci. 2021, 15, 1–19. [Google Scholar] [CrossRef]
  3. Wei, C.; Rao, R.R.; Peng, J.; Huang, B.; Stephens, I.E.L.; Risch, M.; Xu, Z.J.; Shao-Horn, Y. Recommended practices and benchmark activity for hydrogen and oxygen electrocatalysis in water splitting and fuel cells. Adv. Mater. 2019, 31, 1806296. [Google Scholar] [CrossRef] [PubMed]
  4. Norskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  5. Wang, X.; Li, Z.; Qu, Y.; Yuan, T.; Wang, W.; Wu, Y.; Li, Y. Review of metal catalysts for oxygen reduction reaction: From nanoscale engineering to atomic design. Chem 2019, 5, 1486–1511. [Google Scholar] [CrossRef]
  6. Sui, S.; Wang, X.; Zhou, X.; Su, Y.; Riffat, S.; Liu, C.J. A comprehensive review of Pt electrocatalysts for the oxygen reduction reaction: Nanostructure, activity, mechanism and carbon support in PEM fuel cells. J. Mater. Chem. A 2017, 5, 1808. [Google Scholar] [CrossRef]
  7. Shao, M.; Chang, Q.; Dodelet, J.P.; Chenitz, R. Recent advances in electrocatalysts for oxygen reduction reaction. Chem. Rev. 2016, 116, 3594–3657. [Google Scholar] [CrossRef]
  8. Wu, G.; Zelenay, P. Nanostructured nonprecious metal catalysts for oxygen reduction reaction. Acc. Chem. Res. 2013, 46, 1878–1889. [Google Scholar] [CrossRef]
  9. Goswami, C.; Hazarika, K.K.; Bharali, P. Transition metal oxide nanocatalysts for oxygen reduction reaction. Mater. Sci. Energy Technol. 2018, 1, 117–128. [Google Scholar] [CrossRef]
  10. Yang, C.C.; Zai, S.F.; Zhou, Y.T.; Du, L.; Jiang, Q. Fe3C-Co nanoparticles encapsulated in a hierarchical structure of N-doped carbon as a multifunctional electrocatalyst for ORR, OER, and HER. Adv. Funct. Mater. 2019, 29, 1901949. [Google Scholar] [CrossRef]
  11. Liang, Y.; Li, Y.; Wang, H.; Zhou, J.; Wang, J.; Regier, T.; Dai, H. Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction. Nat. Mater. 2011, 10, 780–786. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, M.; Ge, K.; Wu, Y.; Zhang, B.; Duan, J. Synergistic catalysis of cobalt tetroxide and bamboo-shaped carbon nanotubes doped with nitrogen for oxygen reduction in Zn–air batteries. Inorg. Chem. 2023, 62, 13378–13386. [Google Scholar] [CrossRef] [PubMed]
  13. Khusnuriyalova, A.F.; Caporali, M.; Hey-Hawkins, E.; Sinyashin, O.G.; Yakhvarov, D.G. Preparation of cobalt nanoparticles. Eur. J. Inorg. Chem. 2021, 2021, 3023–3114. [Google Scholar]
  14. Gu, M.; Yao, S.B.; Zhou, S.M. Electrochemical study of cobalt electrocrystallization on glass carbon from sulfate solution. Electrochemistry 2006, 74, 309–314. [Google Scholar] [CrossRef]
  15. Zhang, L.; Liu, B.; Zhang, N.; Ma, M. Electrosynthesis of Co3O4 and Co(OH)2 ultrathin nanosheet arrays for efficient electrocatalytic water splitting in alkaline and neutral media. Nano Res. 2018, 11, 323–333. [Google Scholar] [CrossRef]
  16. Lisnund, S.; Blay, V.; Muamkhunthod, P.; Thunyanon, K.; Pansalee, J.; Monkrathok, J.; Maneechote, P.; Chansaenpak, K.; Pinyou, P. Electrodeposition of cobalt oxides on carbon nanotubes for sensitive bromhexine sensing. Molecules 2022, 27, 4078. [Google Scholar] [CrossRef]
  17. Niveditha, C.V.; Aswini, R.; Jabeen Fatima, M.J.; Ramanarayanan, R.; Pullanjiyot, N.; Swaminathan, S. Feather like highly active Co3O4 electrode for supercapacitor application: A potentiodynamic approach. Mater. Res. Express 2018, 5, 065501. [Google Scholar] [CrossRef]
  18. Ling, T.; Yan, D.Y.; Jiao, Y.; Wang, H.; Zheng, Y.; Zheng, X.; Mao, J.; Du, X.-W.; Hu, Z.; Jaroniec, M.; et al. Engineering surface atomic structure of single-crystal cobalt (II) oxide nanorods for superior electrocatalysis. Nat. Commun. 2016, 7, 12876. [Google Scholar] [CrossRef]
  19. Liang, Y.; Wang, H.; Diao, P.; Chang, W.; Hong, G.; Li, Y.; Gong, M.; Xie, L.; Zhou, J.; Wang, J.; et al. Oxygen reduction electrocatalyst based on strongly coupled cobalt oxide nanocrystals and carbon nanotubes. J. Am. Chem. Soc. 2012, 134, 15849–15857. [Google Scholar] [CrossRef]
  20. Mao, S.; Wen, Z.; Huang, T.; Hou, Y.; Chen, J. High-performance bifunctional electrocatalysts of 3d crumpled graphene cobalt oxide nanohybrids for oxygen reduction and evolution reactions. Energy Environ. Sci. 2014, 7, 609–616. [Google Scholar] [CrossRef]
  21. Zhang, S.; Shang, N.; Gao, S.; Meng, T.; Wang, Z.; Gao, Y.; Wang, C. Ultra dispersed Co supported on nitrogen-doped carbon: An efficient electrocatalyst for oxygen reduction reaction and Zn-air battery. Chem. Eng. Sci. 2021, 234, 116442. [Google Scholar] [CrossRef]
  22. Huang, D.; Luo, Y.; Li, S.; Zhang, B.; Shen, Y.; Wang, M. Active catalysts based on cobalt oxide@cobalt/N-C nanocomposites for oxygen reduction reaction in alkaline solutions. Nano Res. 2014, 7, 1054–1064. [Google Scholar] [CrossRef]
  23. Liu, L.; Liu, H.; Sun, X.; Li, C.; Bai, J. Efficient electrocatalyst of PteFe/CNFs for oxygen reduction reaction in alkaline media. Int. J. Hydrogen Energy 2020, 45, 15112–15120. [Google Scholar] [CrossRef]
  24. Alemany-Molina, G.; Lo Vecchio, C.; Baglio, V.; Aricò, A.S.; Morallón, E.; Cazorla-Amorós, D. Pt nanoparticles for improving the performance and durability of Fe-N-C based materials towards oxygen reduction reaction in alkaline direct methanol fuel cells. J. Colloid Sci. 2025, 691, 137426. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, R.; Wang, K.; Wang, P.; He, Y.; Liu, Z. Pt-Fe-Co Ternary metal single atom catalyst for toward high efficiency alkaline oxygen reduction reaction. Energies 2023, 16, 3684. [Google Scholar] [CrossRef]
  26. Campos-Roldán, C.A.; Calvillo, L.; Granozzi, G.; Alonso-Vante, N. Alkaline hydrogen electrode and oxygen reduction reaction on PtxNi nanoalloys. J. Electroanal. Chem. 2020, 857, 113449. [Google Scholar] [CrossRef]
  27. Zysler, M.; Shokhen, V.; Hardisty, S.S.; Muzikansky, A.; Zitoun, D. Bifunctional Pt−Ni electrocatalyst synthesis with ultralow platinum seeds for oxygen evolution and reduction in alkaline medium. ACS Appl. Energy Mater. 2022, 5, 4212–4220. [Google Scholar] [CrossRef]
  28. Liao, Q.; Kuroki, H.; Tamaki, T.; Arao, M.; Matsumoto, M.; Imai, H.; Yamaguchi, T. Three-dimensionally connected platinum-cobalt nanoparticles as support-free electrocatalysts for oxygen reduction. ACS Appl. Nano Mater. 2025, 8, 3323–3332. [Google Scholar] [CrossRef]
  29. Zeng, X.; Mitra, S.K.; Li, X. One-pot synthesis of supported PtCox bifunctional catalysts for oxygen reduction and hydrogen evolution reactions. Int. J. Hydrogen Energy 2024, 86, 577–585. [Google Scholar] [CrossRef]
  30. Yan, W.; Sun, P.; Luo, C.; Xia, X.; Liu, Z.; Zhao, Y.; Zhang, S.; Sun, L.; Du, F. PtCo-based nanocatalyst for oxygen reduction reaction: Recent highlights on synthesis strategy and catalytic mechanism. Chin. J. Chem. Eng. 2023, 53, 101–123. [Google Scholar] [CrossRef]
  31. Oezaslan, M.; Hasché, F.; Strasser, P. Oxygen electroreduction on PtCo3, PtCo and Pt3Co alloy nanoparticles for alkaline and acidic PEM fuel cells. J. Electrochem. Soc. 2012, 159, B394–B405. [Google Scholar] [CrossRef]
  32. Lima, F.H.B.; de Castro, J.F.R.; Santos, L.G.R.A.; Ticianelli, E.A. Electrocatalysis of oxygen reduction on carbon-supported Pt–Co nanoparticles with low Pt content. J. Power Sources 2009, 190, 293–300. [Google Scholar] [CrossRef]
  33. Miyatake, K.; Shimizu, Y. Pt/Co Alloy Nanoparticles Prepared by Nanocapsule Method Exhibit a High Oxygen Reduction Reaction Activity in the Alkaline Media. ACS Omega 2017, 2, 2085–2089. [Google Scholar] [CrossRef] [PubMed]
  34. Sravani, B.; Raghavendra, P.; Chandrasekhar, Y.; Veera Manohara Reddy, Y.; Sivasubramanian, R.; Venkateswarlu, K.; Madhavi, G.; Subramanyam Sarma, L. Immobilization of platinum-cobalt and platinum-nickel bimetallic nanoparticles on pomegranate peel extract-treated reduced graphene oxide as electrocatalysts for oxygen reduction reaction. Int. J. Hydrogen Energy 2020, 45, 7680–7690. [Google Scholar] [CrossRef]
  35. Yun, M.; Ahmed, M.S.; Han, H.S.; Jeon, S. Platinum-cobalt binary alloyed nanoparticles supported on thiolated graphene oxide for oxygen reduction reaction in alkaline media. J. Nanosci. Nanotechnol. 2016, 16, 9675–9682. [Google Scholar] [CrossRef]
  36. Rakočević, L.; Golubović, J.; Vasiljević Radović, D.; Rajić, V.; Štrbac, S. A comparative study of hydrogen evolution on Pt/GC and Pt/GNPs in acid solution. Int. J. Hydrogen Energy 2024, 51, 1240–1254. [Google Scholar] [CrossRef]
  37. Vanysek, P. Electrochemical series. In CRC Handbook of Chemistry and Physics, 93rd ed.; Haynes, W.M., Ed.; CRC Press Taylor & Francis Group: Boca Raton, FL, USA; London, UK, 2012; pp. 8–21. [Google Scholar]
  38. Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous Solutions, 2nd ed.; National Association of Corrosion Engineers: Houston, TX, USA, 1974; p. 325. (In English) [Google Scholar]
  39. Xie, R.-C.; Batchelor-McAuley, C.; Rauwel, E.; Rauwel, P.; Compton, R.G. Electrochemical characterisation of Co@Co(OH)2 core-shell nanoparticles and their aggregation in solution. ChemElectroChem 2020, 7, 4259–4268. [Google Scholar] [CrossRef]
  40. Gomez Meier, H.; Vilche, J.R.; Arvía, A.J. The electrochemical behaviour of cobalt in alkaline solutions Part I. The potentiodynamic response in the potential region of the Co/CoO couple. J. Electroanal. Chem. Interf. Electrochem. 1982, 134, 251–272. [Google Scholar] [CrossRef]
  41. Gomez Meier, H.; Vilche, J.R.; Arvía, A.J. The electrochemical behaviour of cobalt in alkaline solutions part II. The potentiodynamic response of Co(OH)2 electrodes. J. Electroanal. Chem. Interf. Electrochem. 1982, 138, 367–379. [Google Scholar] [CrossRef]
  42. Burke, L.D.; Murphy, M.M. The electrocatalytic behavior of cobalt (and iron) electrodes at low potential in base. J. Electrochem. Soc. 1991, 138, 88–94. [Google Scholar] [CrossRef]
  43. Veggetti, E.; Kodintsev, I.M.; Trasatti, S. Hydrogen evolution on oxide electrodes: Co3O4 in alkaline solution. J. Electroanal. Chem. 1992, 339, 255–268. [Google Scholar] [CrossRef]
  44. Štrbac, S. The effect of pH on oxygen and hydrogen peroxide reduction on polycrystalline Pt electrode. Electrochim. Acta 2011, 56, 1597–1604. [Google Scholar] [CrossRef]
  45. Saeki, R.; Ohgai, T. Effect of growth rate on the crystal orientation and magnetization performance of cobalt nanocrystal arrays electrodeposited from aqueous solution. Nanomaterials 2018, 8, 566. [Google Scholar] [CrossRef]
  46. Manjunatha, M.; Reddy, G.S.; Mallikarjunaiah, K.J.; Ramakrishna, D.; Ramesh, K.P. Determination of phase composition of cobalt nanoparticles using 59Co internal field nuclear magnetic resonance. J. Supercond. Nov. Magn. 2019, 32, 3201–3209. [Google Scholar] [CrossRef]
  47. de la Pena O’Shea, V.A.; Piscina, P.R.; Homs, N.; Aromí, G.; Fierro, J.L.G. Development of hexagonal closed-packed cobalt nanoparticles stable at high temperature. Chem. Mater. 2009, 21, 5637–5643. [Google Scholar] [CrossRef]
  48. Meng, Q.; Guo, S.; Zhao, X.; Veintemillas-Verdaguer, S. Bulk metastable cobalt in fcc crystal structure. J. Alloys Comp. 2013, 580, 187–190. [Google Scholar] [CrossRef]
  49. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  50. Rajagopal, V.; Mehla, S.; Jones, L.A.; Bhargava, S.K. Nanoengineered cobalt electrocatalyst for alkaline oxygen evolution reaction. Nanomaterials 2024, 14, 946. [Google Scholar] [CrossRef]
  51. Milikić, J.; Knežević, S.; Ognjanović, M.; Stanković, D.; Rakočević, L.; Šljukić, B. Template-based synthesis of Co3O4 and Co3O4/SnO2 bifunctional catalysts with enhanced electrocatalytic properties for reversible oxygen evolution and reduction reaction. Int. J. Hydrogen Energy 2023, 48, 27568–27581. [Google Scholar] [CrossRef]
  52. Paul, A.; Gusmão, F.; Mahmoud, A.G.; Hazra, S.; Rakočević, L.; Šljukić, B.; Khan, R.A.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Catalyzing towards clean energy: Tuning the oxygen evolution reaction by amide-functionalized Co(II) and Ni(II) pristine coordination polymers. CrystEngComm 2024, 26, 2755–2764. [Google Scholar] [CrossRef]
  53. Rakočević, L.; Simatović Stojković, I.; Maksić, A.; Rajić, V.; Štrbac, S.; Srejić, I. PtAu nanoparticles supported by reduced graphene oxide as a highly active catalyst for hydrogen evolution. Catalysts 2022, 12, 43. [Google Scholar] [CrossRef]
  54. Štrbac, S.; Adzić, R.R. The influence of OH chemisorption on the catalytic properties of gold single crystal surfaces for oxygen reduction in alkaline solutions. J. Electroanal. Chem. 1996, 403, 169–181. [Google Scholar] [CrossRef]
  55. McCrory, C.C.L.; Jung, S.; Peters, J.C.; Jaramillo, T.F. Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction. J. Am. Chem. Soc. 2013, 135, 16977–16987. [Google Scholar] [CrossRef]
  56. Markovic, N.; Gasteiger, H.; Ross, P.N. Kinetics of oxygen reduction on Pt(hkl) electrodes: Implications for the crystallite size effect with supported Pt electrocatalysts. J. Electrochem. Soc. 1997, 144, 1591. [Google Scholar] [CrossRef]
  57. Siddika, M.; Hosen, N.; Althomali, R.H.; Al-Humaidi, J.Y.; Rahman, M.M.; Hasnat, M.A. Kinetics of electrocatalytic oxygen reduction reaction over an activated glassy carbon electrode in an alkaline medium. Catalysts 2024, 14, 164. [Google Scholar] [CrossRef]
  58. Xiao, J.; Kuang, Q.; Yang, S.; Xiao, F.; Wang, S.; Guo, L. Surface structure dependent electrocatalytic activity of Co3O4 anchored on graphene sheets toward oxygen reduction reaction. Sci. Rep. 2013, 3, 2300. [Google Scholar] [CrossRef]
  59. Zhan, Y.; Lu, M.; Yang, S.; Xu, C.; Liu, Z.; Lee, J.Y. Activity of transition-metal (manganese, iron, cobalt, and nickel) phosphates for oxygen electrocatalysis in alkaline solution. ChemCatChem 2016, 8, 372–379. [Google Scholar] [CrossRef]
  60. Huang, K.; Xu, P.; He, X.; Wang, R.; Wang, Y.; Yang, H.; Zhang, R.; Lei, M.; Tang, H. Annealing-free platinum cobalt alloy nanoparticles on nitrogen-doped mesoporous carbon with boosted oxygen electroreduction performance. ChemElectroChem 2020, 7, 3341–3346. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Liu, J.; Zhao, Y.; Wang, F. Composition-controlled synthesis of carbon-supported Pt–Co alloy nanoparticles and the origin of their ORR activity enhancement. Phys. F Chem. Chem. Phys. 2014, 16, 19298–19306. [Google Scholar] [CrossRef]
  62. Todoroki, N.; Wadayama, T. Oxygen reduction and oxygen evolution reaction activity on Co/Pt(111) surfaces in alkaline solution. ECS Trans. 2018, 86, 569. [Google Scholar] [CrossRef]
  63. Hu, S.; Goenaga, G.; Melton, C.; Zawodzinski, T.A.; Mukherjee, D. PtCo/CoOx nanocomposites: Bifunctional electrocatalysts for oxygen reduction and evolution reactions synthesized via tandem laser ablation synthesis in solution-galvanic replacement reactions. Appl. Catal. B Environ. 2016, 182, 286–296. [Google Scholar] [CrossRef]
  64. Ribeiro, E.L.; Davis, E.M.; Mokhtarnejad, M.; Hu, S.; Mukherjee, D.; Khomami, B. MOF-derived PtCo/Co3O4 nanocomposites in carbonaceous matrices as high-performance ORR electrocatalysts synthesized via laser ablation techniques. Catal. Sci. Technol. 2021, 11, 3002–3013. [Google Scholar] [CrossRef]
  65. Ni, C.; Chen, X.; Chen, Y.; Li, S.; Zhou, T.; Yang, J.; Liu, M.; Su, H. Ultrafine intermetallic platinum-cobalt with a contracted Pt–Pt pair for efficient acidic oxygen reduction reactions. Nanoscale 2025, 17, 10380–10388. [Google Scholar] [CrossRef] [PubMed]
  66. Lu, B.; Liu, Q.; Nichols, F.; Mercado, R.; Morris, D.; Li, N.; Zhang, P.; Gao, P.; Ping, Y.; Chen, S. Oxygen reduction reaction catalyzed by carbon-supported platinum few-atom clusters: Significant enhancement by doping of atomic cobalt. Research 2020, 2020, 9167829. [Google Scholar] [CrossRef] [PubMed]
  67. Huang, L.; Wei, M.; Qi, R.; Dong, C.-L.; Dang, D.; Yang, C.-C.; Xia, C.; Chen, C.; Zaman, S.; Li, F.-M.; et al. An integrated platinum-nanocarbon electrocatalyst for efficient oxygen reduction. Nat. Commun. 2022, 13, 6703. [Google Scholar] [CrossRef]
  68. Lee, Y.; Lee, J.H.; Lee, D.; Oh, S.; Park, J.; Im, K.; Yoo, S.J.; Kim, J. Low-loading platinum–cobalt electrocatalyst supported on hollow carbon for enhanced oxygen reduction reaction. Chem. Eng. J. 2024, 500, 157072. [Google Scholar] [CrossRef]
  69. Li, F.; Liu, Y.; Dong, Y.; Wang, H.; Wang, T.; Lu, J.; Huang, W.; Chen, X.; Guo, Y.; Zheng, X.; et al. Electronic modulation of PtCo catalysts by topological carbon defects boosting durable oxygen reduction reaction in acidic media. Fundam. Res. 2025; in press. [Google Scholar] [CrossRef]
  70. Holade, Y.; Sahin, N.E.; Servat, K.; Napporn, T.W.; Kokoh, K.B. Recent advances in carbon supported metal nanoparticles preparation for oxygen reduction reaction in low temperature fuel cells. Catalysts 2015, 5, 310–348. [Google Scholar] [CrossRef]
Figure 1. Electrochemical characterization of Cocryst/GC and Cocryst-Pt/GC electrodes in 0.1 M NaOH: (a) CVs recorded after a 30 min deposition of Co at a potential of −0.36 V; (b) OCP curve recorded after immersion of the obtained Cocryst/GC electrode into the Pt depositing solution; (c) CV of the Cocryst-Pt/GC obtained after 60 min spontaneous Pt deposition on Cocryst/GC; (d) CVs of Ptspont/GC, obtained after 60 min spontaneous Pt deposition on GC, and bare Pt(poly) electrodes. The scan rate was 50 mV/s.
Figure 1. Electrochemical characterization of Cocryst/GC and Cocryst-Pt/GC electrodes in 0.1 M NaOH: (a) CVs recorded after a 30 min deposition of Co at a potential of −0.36 V; (b) OCP curve recorded after immersion of the obtained Cocryst/GC electrode into the Pt depositing solution; (c) CV of the Cocryst-Pt/GC obtained after 60 min spontaneous Pt deposition on Cocryst/GC; (d) CVs of Ptspont/GC, obtained after 60 min spontaneous Pt deposition on GC, and bare Pt(poly) electrodes. The scan rate was 50 mV/s.
Catalysts 15 00490 g001
Figure 2. XRD and SEM characterization of Cocryst/GC (left column) and Cocryst-Pt/GC (right column): (a,b) XRD patterns; (c,d) SEM images (magnification = 35,000×, scale bar 1 μm); (e,f) size distributions of Cocryst and Cocryst-Pt islands.
Figure 2. XRD and SEM characterization of Cocryst/GC (left column) and Cocryst-Pt/GC (right column): (a,b) XRD patterns; (c,d) SEM images (magnification = 35,000×, scale bar 1 μm); (e,f) size distributions of Cocryst and Cocryst-Pt islands.
Catalysts 15 00490 g002
Figure 3. Elemental mapping and EDX analysis of the individual Cocryst (left column) and Cocryst-Pt islands (right column): (a,b) the magnification of the SEM image of the Cocryst island is 120,000×, and of the Cocryst-Pt island is 100,000×; (c,d) corresponding EDX chemical analysis.
Figure 3. Elemental mapping and EDX analysis of the individual Cocryst (left column) and Cocryst-Pt islands (right column): (a,b) the magnification of the SEM image of the Cocryst island is 120,000×, and of the Cocryst-Pt island is 100,000×; (c,d) corresponding EDX chemical analysis.
Catalysts 15 00490 g003
Figure 4. XPS characterization of Cocryst/GC (left column) and Cocryst-Pt/GC (right column): (a) survey spectrum of Cocryst/GC (b) high-resolution spectra of Co 2p, and (c) O 1s for Cocryst/GC; (d) survey spectrum of Cocryst-Pt/GC; (e) high-resolution spectra of Pt 4f7/2, and (f) O 1s for Cocryst-Pt/GC.
Figure 4. XPS characterization of Cocryst/GC (left column) and Cocryst-Pt/GC (right column): (a) survey spectrum of Cocryst/GC (b) high-resolution spectra of Co 2p, and (c) O 1s for Cocryst/GC; (d) survey spectrum of Cocryst-Pt/GC; (e) high-resolution spectra of Pt 4f7/2, and (f) O 1s for Cocryst-Pt/GC.
Catalysts 15 00490 g004
Figure 5. Polarization curves for ORR in 0.1 M NaOH on: (a) Cocryst/GC; (b) corresponding K–L plot; (c) Pt-Cocryst, and (d) corresponding K–L plot. The scan rate was 50 mV s−1.
Figure 5. Polarization curves for ORR in 0.1 M NaOH on: (a) Cocryst/GC; (b) corresponding K–L plot; (c) Pt-Cocryst, and (d) corresponding K–L plot. The scan rate was 50 mV s−1.
Catalysts 15 00490 g005
Figure 6. Comparison of the ORR activity of the GC-supported Ptspont, Cocryst, and Cocryst-Pt catalysts and Pt(poly): (a) LSV curves recorded in oxygen-saturated 0.1 M NaOH with a scan rate of 50 mV s−1; (b) Corresponding Tafel plots.
Figure 6. Comparison of the ORR activity of the GC-supported Ptspont, Cocryst, and Cocryst-Pt catalysts and Pt(poly): (a) LSV curves recorded in oxygen-saturated 0.1 M NaOH with a scan rate of 50 mV s−1; (b) Corresponding Tafel plots.
Catalysts 15 00490 g006
Figure 7. Long-term stability testing of Cocryst-Pt catalyst during ORR in oxygen-saturated 0.1 M NaOH: (a) CA curves recorded over 5 h at a constant applied potential as indicated and a rotation rate of 1600 rpm. CA curves for Cocryst alone and Pt(poly) are presented for comparison. (b) LSV curves recorded before and after 5 h stability testing. (c) LSV curves recorded before and after 2000 cycles. LSV curves were recorded with a scan rate of 50 mV s−1.
Figure 7. Long-term stability testing of Cocryst-Pt catalyst during ORR in oxygen-saturated 0.1 M NaOH: (a) CA curves recorded over 5 h at a constant applied potential as indicated and a rotation rate of 1600 rpm. CA curves for Cocryst alone and Pt(poly) are presented for comparison. (b) LSV curves recorded before and after 5 h stability testing. (c) LSV curves recorded before and after 2000 cycles. LSV curves were recorded with a scan rate of 50 mV s−1.
Catalysts 15 00490 g007
Table 1. Atomic and weight percentage of the elements for Cocryst/GC and Cocryst-Pt/GC.
Table 1. Atomic and weight percentage of the elements for Cocryst/GC and Cocryst-Pt/GC.
LineCocryst/GCCocryst-Pt/GC
at%wt%at%wt%
C 1s11.66.550.524.7
O 1s74.455.236.223.6
Co 2p14.038.3--
Pt 4f7/2--5.039.7
Cl 2p--8.312.0
Table 2. Comparison of the ORR activity of various Pt-Co catalysts.
Table 2. Comparison of the ORR activity of various Pt-Co catalysts.
CatalystSolutionEonset (V) E1/2 (V)Tafel Slope
(mV dec−1)
Reference
Pt3Co/r-GO0.1 M KOH1.130.76-[34]
ER/PtCo-tG0.1 M KOH0.940.86-[35]
PtCo@NMC0.1 M KOH1.040.9667.4[60]
Pt76Co241 M NaOH0.98-59.98[61]
Co/Pt(111)0.1 M KOH--51[62]
PtCo-31 M KOH-0.8639.2[63]
125-PtCo@C–Co3O41 M KOH0.95-38[64]
Pt3Co/NC1 M KOH-0.91-[65]
PtCo-NC-40.1 M HClO41.030.9353.7[66]
PtCo@CoNC/NTG0.1 M HClO4-0.9471[67]
5-wt% Pt/h-Co-NC0.1 M HClO41.000.8791[68]
PtCo-DPC0.1 M HClO41.020.8587.5[69]
Cocryst-Pt0.1 M NaOH1.070.8763This work
Pt/C0.1 M KOH1.010.89102.6[60]
20 wt%Pt/C0.1 M NaOH-0.8569[70]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Golubović, J.; Rakočević, L.; Rajić, V.; Milović, M.; Štrbac, S. Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction. Catalysts 2025, 15, 490. https://doi.org/10.3390/catal15050490

AMA Style

Golubović J, Rakočević L, Rajić V, Milović M, Štrbac S. Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction. Catalysts. 2025; 15(5):490. https://doi.org/10.3390/catal15050490

Chicago/Turabian Style

Golubović, Jelena, Lazar Rakočević, Vladimir Rajić, Miloš Milović, and Svetlana Štrbac. 2025. "Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction" Catalysts 15, no. 5: 490. https://doi.org/10.3390/catal15050490

APA Style

Golubović, J., Rakočević, L., Rajić, V., Milović, M., & Štrbac, S. (2025). Electrodeposited Co Crystalline Islands Shelled with Facile Spontaneously Deposited Pt for Improved Oxygen Reduction. Catalysts, 15(5), 490. https://doi.org/10.3390/catal15050490

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop