Chih-Chun
Chang
a,
Jui-Cheng
Kao
b,
Yu-Chieh
Lo
*b,
Jyh-Pin
Chou
*c,
Shang-Cheng
Lin
d,
Chun-Chia
Wen
a and
Michael H.
Huang
*a
aDepartment of Chemistry, National Tsing Hua University, Hsinchu 300044, Taiwan. E-mail: hyhuang@mx.nthu.edu.tw
bDepartment of Materials Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan. E-mail: yclo@nycu.edu.tw
cGraduate School of Advanced Technology, National Taiwan University, Taipei 106319, Taiwan. E-mail: jpchou@ntu.edu.tw
dDepartment of Chemical Engineering, National Tsing Hua University, Hsinchu 300044, Taiwan
First published on 20th March 2025
Examination of the facet effects of metal oxide crystals on the oxygen reduction reaction (ORR) has been inadequately investigated due to the limited availability of polyhedra that expose only specific surfaces. Here, cuprous oxide cubes, octahedra, and rhombic dodecahedra, exposing the respective {100}, {111}, and {110} surfaces, were incorporated into a matrix of carbon nanotubes (CNTs) to enhance electrical conductivity. The composites were evaluated for their electrocatalytic ORR activity. The rhombic dodecahedra/CNTs composite exhibited the highest ORR activity, followed by the octahedra/CNTs and then the cubes/CNTs. Commercial Cu2O powder/CNTs showed notably lower ORR activity, demonstrating the importance of catalyst surface control for ORR performance. Koutecký–Levich plots showed that these Cu2O polyhedra were highly selective towards the four-electron pathway in the ORR, whereas the commercial Cu2O powder/CNTs catalyst proceeded via the two-electron pathway. Durability tests revealed a reversed trend, with the cubes/CNTs being the most stable electrocatalyst. Density functional theory (DFT) calculations indicated the weakest O2 adsorption on the Cu2O {110} surface. The free energy diagrams and 2D volcano plot further established the {110} surface as the most active towards ORR, while strong OH intermediate binding on the {100} and {111} surfaces led to lower theoretical limiting potentials and higher overpotentials. DFT results provided mechanistic insights to explain the experimental facet effects.
Considering the high cost of platinum-based electrocatalysts with superior ORR performance, transition metal oxide electrocatalysts have been explored as cost-effective and highly active alternatives.15,16 To improve the electrical conductivity of transition metal oxide electrocatalysts, decoration or dispersion with metal nanoparticles, carbon-based materials, and conducting polymers is effective in enhancing interparticle conductivity and catalytic activity.17–19 With regard to ORR activity, the exposed crystal faces should be a significant factor, especially considering the existence of face-related lattice variations in semiconductor crystals. Previously, Co3O4 spheres, octahedra, and truncated octahedra on reduced graphene oxide sheets were evaluated for ORR activity, with the octahedra showing the highest activity.20 Carbon-dispersed Cu2O nanocubes, truncated cubes, and nanoporous particles were also compared for ORR performance.21 However, it was not demonstrated that the polymers added to synthesize the Cu2O crystals were completely removed. These Cu2O particle morphologies, prepared by electrochemical deposition, were also examined for their ORR activities. However, the obtained polarization curves did not exhibit limiting currents.22 Au particle-decorated Cu2O spheres, cubes, and multipods dispersed onto a carbon nanotube matrix were also employed for the oxygen reduction reaction.23 However, a comprehensive understanding of the facet effects on the ORR activity of Cu2O crystals is still lacking.
In this work, Cu2O cubes, octahedra, and rhombic dodecahedra, exposing the {100}, {111}, and {110} faces, respectively, were mixed with a carbon nanotube matrix to form active electrocatalysts for the oxygen reduction reaction. Their ORR linear sweep voltammetry (LSV) polarization curves, after normalization for active surface area, were compared. The number of transferred electrons was determined for each sample. Chronoamperometric curves were obtained. Moreover, DFT calculations of oxygen adsorption energy on the three Cu2O surfaces and the free energy changes associated with intermediate formation were performed to support the experimental results. An ORR activity map was also constructed.
Initially, the electrochemical performance of these Cu2O crystals was evaluated by determining the redox peak reversibility through the cyclic voltammetry curves of potassium ferricyanide (K3[Fe(CN)6]). Fig. S4 (ESI)† presents the CV curves. The redox peaks of [Fe(CN)6]3−/4− measured using these Cu2O crystals did not show good reversibility, with less clear peak positions and low current densities. The octahedra delivered the largest current density, followed by the cubes and rhombic dodecahedra. This order was the same as the order of their electrical conductivity behaviors, so it is clear that their intrinsic electrical conductivity properties greatly affected the electrochemical performance.
To address the low electrochemical activity issue, electroconductive carbon black and carbon nanotubes were introduced to improve the overall stability and activity of the catalysts. Fig. 1 gives the SEM images of the Cu2O cubes, octahedra, and rhombic dodecahedra, showing they were well dispersed in the CNT matrix. The Cu2O particles still retained their shapes. The XRD patterns of the composites showed a diffraction peak from the CNTs (Fig. S5, ESI†). Interestingly, the diffraction peaks for the octahedra were shifted to slightly higher 2θ angles than those for the rhombic dodecahedra, so shape-related lattice constant changes could still be observed. Fig. S6 (ESI)† displays the measured cyclic voltammetry (CV) curves. The Cu2O crystals mixed with carbon black and CNTs could all greatly improve the current density and reversibility of the [Fe(CN)6]3−/4− redox peaks, and mixing the crystals with CNTs produced higher current densities. Thus, the Cu2O/CNT composite was used for the subsequent electroanalytical measurements.
![]() | ||
Fig. 1 SEM images of Cu2O (a) cubes, (b) octahedra, and (c) rhombic dodecahedra mixed with carbon nanotubes. |
Electrochemical impedance spectroscopy (EIS) measurements were then performed to examine the impedance behavior of these catalysts. The Nyquist plots were fitted using a Randles circuit model to obtain the impedance values. Previously, the charge-transfer resistance (Rct) values of the Cu2O cubes, cuboctahedra, and rhombic dodecahedra without mixing with CNTs were in the range of 18.9 to 28.9 kΩ.3 As seen in Fig. 2, after mixing with the CNTs, all the Cu2O catalysts exhibited very small impedance values. The trend for charge transfer was CNTs > octahedra/CNTs > cubes/CNTs > rhombic dodecahedra/CNTs. The Rct values for the Cu2O/CNTs were all approximately <10 Ω. Since the Rct values were too small, the fitting results could only be estimated. By mixing with CNTs, the overall electrical conductivity and charge-transfer ability of the catalysts were greatly enhanced, rendering them more suitable for electrochemical experiments.
Before conducting electrocatalytic ORR measurements with LSV and chronoamperometric analysis (CA) to evaluate the performance and stability of the Cu2O/CNTs catalysts, the variation in the electrochemically active surface area of each sample needed to be considered. When comparing the catalytic performance of the catalysts, the measured current densities of the LSV curves had to account for differences in electrochemically active surface area (ECSA) to provide a more accurate evaluation of the facet effects on ORR activity. The double-layer capacitance (Cdl) was used to estimate the corresponding ECSA differences. The double-layer capacitance was determined from the CV curves using the following equation:
Δj(ja − jc)/2 = Cdl × v | (1) |
ECSA = Cdl/Cs | (2) |
The corresponding CV curves and a plot of charging current density Δj(ja − jc) vs. scan rate (v) are shown in Fig. S7 (ESI).† The Cdl values of the rhombic dodecahedra/CNTs, cubes/CNTs, octahedra/CNTs, commercial Cu2O powder/CNTs, and CNTs were 7.28, 6.46, 6.32, 5.70, and 5.30 mF cm−2, respectively (see Table S2, ESI†). To eliminate the ECSA effect, the ORR LSV curves were normalized using the following equation:
Normalized current density = current density/Cdl ratio | (3) |
O2 + 2H2O + 4e− ⇌ 4OH−; E° = 0.401 V vs. SHE | (4) |
For the alternative two-electron pathway, O2 is first reduced to peroxide ion and can be further reduced (eqn (5) and (6)).
O2 + H2O + 2e− ⇌ HO2− + OH−; E° = −0.076 V vs. SHE | (5) |
H2O + HO2− + 2e− ⇌ 3OH−; E° = 0.878 V vs. SHE | (6) |
ERHE = ESHE + 0.059 × pH | (7) |
Fig. 3 shows the ORR steady-state linear sweep voltammetry polarization curves for different electrocatalysts in an O2-saturated KOH electrolyte, with an oxygen flow at a scan rate of 5 mV s−1 and an electrode rotating rate of 1600 rpm. The expected ORR limiting currents were observed for all the samples, showing the positive effect of stability and performance with the addition of CNTs. Compared with the CNTs and the commercial Cu2O powder, which have large particle sizes and random morphologies, the Cu2O polyhedra/CNTs exhibited notably better ORR catalytic activities, with the rhombic dodecahedra showing the largest limiting current density of −4.25 mA cm−2, followed by the octahedra at −3.93 mA cm−2 and the cubes at −3.67 mA cm−2. This highlights the importance of catalyst surface control on catalytic activity.
Next, a series of linear sweep voltammetry polarization curves for the cubes, octahedra, and rhombic dodecahedra mixed with CNTs were obtained at a scan rate of 5 mV s−1 and different rotating rates (i.e., 400, 600, 900, 1200, 1600, 2000, 2500, and 3600 rpm), as shown in Fig. 4. The data were recorded after scanning multiple times until the curves became stabilized, ensuring consistent and stable performance of the catalysts. These data were normalized with the ECSA for a better facet-related electrocatalytic activity comparison. It was found that the catalytic current increased progressively as the rotating rate increased. The rhombic dodecahedra/CNTs catalyst remained the most electrocatalytically active in terms of ORR performance, with current densities of −5.09 mA cm−2 at 1600 rpm and −10.4 mA cm−2 at 3600 rpm. The octahedra/CNTs also achieved a respectable limiting current density of −4.44 mA cm−2 at 1600 rpm, while it was −3.81 mA cm−2 at 1600 rpm for the cubes/CNTs (see Fig. S8, ESI†). This notable facet dependence in ORR activity may be due to the different interactions between the crystal facets and oxygen.
To further understand the ORR catalytic mechanism and kinetics, the LSV polarization curves at different rotating rates were used to obtain the Koutecký–Levich plots for the different potentials. The Koutecký–Levich plot is used in the study of electrode reaction kinetics and the evaluation of catalyst activity. As can be seen in Fig. 4, the corresponding Koutecký–Levich plots exhibited good linearity in the range of mass-transport control from 0.35 V to 0.5 V, demonstrating a first-order reaction characteristic with oxygen.
Next, the average number of electrons transferred (n) during the ORR was extracted from the Koutecký–Levich plot, which helps in determining whether the oxygen reduction reaction follows a four-electron or a two-electron reaction pathway through the Koutecký–Levich equation (eqn (8)).
1/J = 1/JL + 1/JK = 1/Bω1/2 + 1/JK | (8) |
B = 0.62nFADO2/3v−1/6CO | (9) |
The stability of the catalysts was evaluated by chronoamperometric (i–t) measurements for 6 h under a constant potential of 0.65 V vs. RHE at a rotating rate of 1600 rpm in O2-saturated 0.1 M KOH solution. Fig. 5 shows that the cubes/CNTs sample retained 89.9% of its original current density after 6 h, showing it had the best stability. Meanwhile, the octahedra/CNTs catalyst retained 82.7% of its original current density, and the rhombic dodecahedra/CNTs catalyst displayed the fastest decay rate, retaining just 75.4% of its original current density. The gradual decrease in current density implied that the effect of any deposited Pt clusters on the working electrode was not a concern for this work. The CA measurements clearly showed that mixing with CNTs could improve the catalyst stability for the ORR. Mechanistically, the inferior stability of the rhombic dodecahedra/CNTs could be attributed to the {110} crystal surface of Cu2O being more reactive with oxygen and self-oxidizing to CuO during the ORR. As the crystal surface changed, the current density in the CA curve gradually decreased. This further indicated that the rhombic dodecahedra/CNTs composite exhibited superior electrocatalytic performance and ORR catalytic activity compared to the other samples. Fig. S10 (ESI)† shows the SEM images of the Cu2O crystals after the chronoamperometric measurements. Their morphologies had become less recognizable. However, the cubes maintained a better morphology than the octahedra and rhombic dodecahedra, so the cubes had better stability but lower reactivity. The X-ray fluorescence spectroscopic measurements showed no presence of Pt on the Cu2O cubes/CNTs electrode after the chronoamperometric experiment (Fig. S11†). Fig. S12 (ESI)† provides SEM images of the three Cu2O/CNTs samples after storing them for two months in sealed vials containing isopropanol. The cubes and octahedra still maintained their original morphologies, while the rhombic dodecahedra had been etched due to a possible reaction with trace oxygen in the isopropanol.
![]() | ||
Fig. 5 Chronoamperometric curves for the Cu2O/CNTs catalysts measured at 0.65 V vs. RHE in an O2-saturated 0.1 M KOH solution. |
Tafel plots offer the electrochemical kinetics relationship between the current (or reaction rate) and overpotential through the Tafel equation (eqn (10)).
η = a − b log i | (10) |
Under the cathodic reduction reaction, η is the overpotential (V), i is the current density (A m−2), a is a constant related to the magnitude of the forward and reverse current at equilibrium, and b is the Tafel slope (mV dec−1). Fig. S13 (ESI)† gives the Tafel plots of the Cu2O cubes, octahedra, and rhombic dodecahedra mixed with CNTs. The Tafel slopes for the cubes/CNTs, octahedra/CNTs, and rhombic dodecahedra/CNTs were 79.1, 75.6, and 73.1 mV dec−1, respectively. These values were lower than the reported Tafel slopes for the Pt/C electrodes (80.4 and 98 mV dec−1),25,26 so the Cu2O/CNTs catalysts could be more efficient in producing current with less overpotential in the electrocatalytic ORR than commercial Pt/C.
The adsorption strength of oxygen is strongly correlated with the oxygen reduction reaction performance. The oxygen adsorption energies (Eads) on the three different Cu2O surfaces were determined using density functional theory calculations. The computational details are provided in the ESI.† We initially investigated the O2 adsorption behavior on the three Cu2O surfaces, and the most stable geometric configurations are displayed in Fig. 6. The Cu–O distances between the O2 and the surface were 3.14, 2.21, and 1.86 Å for the {100}, {110}, and {111} surfaces, respectively. O2 molecules preferentially bound to the Cu atoms on the {110} and {111} surfaces. On the {100} surface, O2 slightly bound to the saturated Cu atoms at a two-fold bridge site. The Eads values of an O2 molecule on the three Cu2O facets followed the order: {110} (−0.09 eV) < {100} (−0.85 eV) < {111} (−1.66 eV). The weakest adsorption strength of oxygen on the {110} surface indicated that a smaller energy barrier was required to form the intermediate product OOH*, leading to the best ORR reactivity. Conversely, the strongest adsorption on the {111} surface could be attributed to the bonding nature of the unsaturated Cu atoms. Herein, the moderate adsorption energy of the {100} surface was mostly contributed by the bond distortion of the surface atoms, as shown in Fig. 6.
![]() | ||
Fig. 6 The most stable O2 adsorption configurations on Cu2O (a) {100}, (b) {110}, and (c) {111} surfaces. The upper and lower panels show the top and side views, respectively. |
The oxygen reduction reaction can involve four elementary steps (see ESI†). The free energy diagrams of the ORR intermediate products (OOH*, O*, OH*) at zero electrode potential (U = 0 V) and equilibrium electrode potential (U = 1.23 V) were evaluated. All the intermediate steps were exothermic at U = 0 V, implying that the ORR pathway was thermodynamically favorable (see Fig. 7a). Here * refers to the active site on the catalyst surface. The free energy of OOH* for the {110} surface (4.63 eV) was higher than that for the {100} (3.57 eV) and {111} (3.47 eV) surfaces, indicating a much weaker OOH* binding ability on the {110} surface. The same scenario could be observed for the O* and OH* intermediates. Significantly, the theoretical limiting potential (UL), or the potential-limiting step, acquired from the minimum change in free energy for each reaction step, indicated that the ORR activity followed the trend {110} (0.29 V) > {100} (0.11 V) > {111} (0.03 V) surfaces. The limiting potential for the {110} surface was the * + O2 + (H+ + e−) → OOH* step, while the limiting potential for the {100} and {111} surfaces was the OH* + (H+ + e−) → * + H2O step. The highest UL and the corresponding best ORR activity of the {110} surface among the three facets aligned well with the experimental observations. The {100} and {111} surfaces showed only a small difference in their ORR performance. In addition to the four-electron process, we also considered the free energy of the two-electron process (red line in Fig. 7a). The free energies of O* (ΔGO*) among these three facets were much lower than that of H2O2, meaning that the oxygen reduction reaction preferred to proceed via the four-electron pathway, consistent with the experimental results.
Fig. 7b presents the free energy diagram at U = 1.23 V. The potential-determining steps (PDSs) and the corresponding overpotentials (ηORR) of the ORR were obtained, where the PDS represented the key step with the maximum free energy rise during the ORR process. The PDS of the {110} surface was the transformation of O2 to OOH* (* + O2 + (H+ + e−) → OOH*), which differed from the other two facets (OH* + (H+ + e−) → * + H2O). The order of the ηORR values was: {110} (0.94 V) < {100} (1.12 V) < {111} (1.20 V). A lower ηORR indicated improved ORR activity. Therefore, the results of the overpotential were consistent with the theoretical limiting potential.
Next, the ORR activity map, as developed by Nørskov et al., is shown in Fig. 8, which presents two important descriptors, namely ΔGOOH* and ΔGOH*.27,28 The best ORR activity was located at ΔGOH* = 1.23 eV and ΔGOOH* = 3.69 eV, corresponding to the highest limiting potential (UL = 1.23 V, see the red hill). Furthermore, the competitive free energies of the ORR steps distinguished the three PDSs into different regimes, which are bounded by the white dashed lines in Fig. 8. The free energies of the intermediates on the Cu2O {100}, {110}, and {111} surfaces were respectively labeled on the 2D volcano plot. Evidently, the {110} surface possessed the best activity among the three Cu2O facets, showing weak OOH* binding on the surface. On the other hand, the {100} and {111} surfaces, with small ΔGOH* values, demonstrated strong OH* binding ability. This scenario could cause catalyst poisoning due to the hard desorption of OH*, resulting in non-ideal ORR activity. The positions of these three facets on the activity map confirmed the key PDSs in the entire ORR pathway, agreeing with the results from the free energy diagram at U = 1.23 V. Besides, these three facets exhibited a linear relationship between ΔGOOH* and ΔGOH*. ΔGOOH* can be expressed as ΔGOOH* = 0.975ΔGOH* + 3.45. This behavior follows the linear scaling relation observed in other transition metal oxides and metal surfaces.29,30 It could be noted that the slope of the scaling line was very close to 1, which implied that OOH* bound to the surface via a single bond.31 The scaling relationship and the descriptor-associated analysis can help explain the catalytic behavior and provide theoretical insight into the performance of the chemical reaction.
![]() | ||
Fig. 8 2D volcano plot for the oxygen reduction reaction. The theoretical limiting potential represents the ORR activity. |
The Hg/HgO electrode potentials could be converted to the reversible hydrogen electrode (RHE) scale according to the Nernst equation:
E (V vs. RHE) = E (V vs. Hg/HgO) + 0.098 + 0.059 × pH | (11) |
E (V vs. RHE) = E (V vs. Ag/AgCl) + 0.197 + 0.059 × pH | (12) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta08855g |
This journal is © The Royal Society of Chemistry 2025 |