Yingjie
He
a,
Drew
Aasen
b,
Haoyang
Yu
a,
Matthew
Labbe
b,
Douglas G.
Ivey
*b and
Jonathan G. C.
Veinot
*a
aDepartment of Chemistry, University of Alberta, 11227 Saskatchewan Drive, Edmonton, Alberta T6G 2G2, Canada. E-mail: jveinot@ualberta.ca
bDepartment of Chemical and Materials Engineering, University of Alberta, 9211 116 St, Edmonton, Alberta T6G 1H9, Canada
First published on 19th June 2020
Hybrids of Mn3O4 nanoparticles and hollow carbon spheres prepared from templated pyrolysis of polydopamine were assembled via a straightforward sonication procedure. The resulting hybrids exhibit excellent catalytic activity toward the oxygen reduction reaction (ORR) in prototype Zn–air batteries. Impressively, these catalysts exhibit higher discharge potential and exceptional stability when compared to commercial Pt–Ru catalysts while simultaneously showing comparable onset potential and maximum current density.
Among the many technologies that have appeared, metal–air batteries have attracted substantial attention in part because of their comparatively high energy density.6–10 Aqueous Zn–air batteries are particularly attractive for stationary energy storage because they are based upon earth abundant, non-toxic elements (i.e., Zn) and have a high theoretical energy density (1086 kW kg−1) which is four times that of current Li-ion batteries.11 Despite these advantages, further advances are required if Zn–air batteries are to realize their full potential. The most substantial benefits will come from the development of efficient precious metal-free catalysts that facilitate reactions at the air electrode.12
The oxygen reduction and evolution reactions (i.e., ORR and OER, respectively) occur at the air electrode and are crucial to battery performance (Fig. 1). When the battery is discharging, the ORR reaction is active and oxygen is reduced to form zincate anions at the Zn electrode. Upon charging, this process is reversed and oxygen is regenerated via the OER. While ORR is thermodynamically favourable under operational conditions, both the ORR and OER are kinetically hindered.13 As a result, electrocatalysts are essential to realizing functional/cyclable batteries. Conventional ORR and OER catalysts rely on precious metals and their oxides.14 The high cost of these metals/metal oxides potentially limits the economic viability of their large-scale application.
Precious-metal free systems are being aggressively explored in efforts to find alternative, efficient catalysts. Of late, a variety of carbon nanomaterials (e.g., N-doped carbon nanotubes (CNTs), carbon nanoribbons, and graphene nanosheets) have shown excellent ORR and/or OER catalytic activity.15–18 Transition metal oxide nanoparticles (e.g., MnOx, CoOx) also exhibit promising performance.19,20 There have even been reports of synergistic effects when these materials are combined, however its chemical origin was not identified.21 Previously we demonstrated that N-doped hollow mesoporous carbon (HMC) nanospheres for their catalytic activity.22
Here, we report a new precious-metal free hybrid that combines the favourable ORR properties of N-doped HMCs with those of earth abundant Mn3O4 nanoparticles. The hybrid exhibits enhanced catalytic ORR performance when compared with the individual components, as well as benchmark Pt–RuO2 catalysts. Further adding to the appeal of this new catalyst, its stability exceeds that of Pt–RuO2 standards when incorporated into prototype Zn–air batteries.
Scanning electron microscopy (SEM) was performed using a Zeiss Sigma 300 VP-FESEM (accelerating voltage of 5–20 kV) equipped with secondary and backscattered electron detectors and an in-lens detector. Samples were prepared by placing a drop of ethanol suspension of the material of interest onto an Al stub and was subsequently air dried for 1 h at room temperature (25 °C). Transmission electron microscopy (TEM) was performed with a JEOL JEM-ARM200CF TEM/STEM (accelerating voltage of 200 kV) equipped with energy dispersive X-ray (EDX) spectrometer. High resolution TEM (HRTEM) images were processed using Gatan Digital Micrograph software (Version 3.22.1461.0) and ImageJ (Version 1.52R). Selected area electron diffraction (SAED) was performed on a JEOL 2010 TEM. TEM samples were prepared by dispersing the purified sample in anhydrous ethanol, which was then drop cast onto a holy/lacey carbon coated Cu grid (Electron Microscopy Inc.) and dried under vacuum for at least 16 h. At least 300 nanoparticles were used for determination of thickness and diameter of the purified materials. EDX maps of several representative regions were also obtained.
X-ray photoelectron spectroscopy (XPS) was performed with a Kratos Axis 165 Ultra X-ray spectrometer operating in energy spectrum mode under ultra-high vacuum. A monochromatic Al K source (λ = 8.34 Å) was used as the X-ray source. The samples were prepared by drop-casting an ethanol suspension of material of interest onto a Cu foil that was air dried for at least 16 h at room temperature (25 °C). The take-off angle was 90°. CasaXPS software (VAMAS) was used to analyse the obtained data. In general, a Shirley-type background was subtracted.24 The spectra were calibrated by setting the deconvoluted adventitious C 1s peak to 284.8 eV.
Powder X-ray diffraction (PXRD) was conducted by placing purified samples on a zero-background Si wafer. PXRD patterns were collected using a Rigaku Ultima IV XRD system equipped with a Cu Kα radiation source.
Nitrogen adsorption–desorption isotherms were obtained with a Quantachrome Autosorb-iQ-XR system at −196 °C. Before measurements, samples were outgassed at 150 °C under vacuum for 3 h. Data was analysed using Brunauer–Emmett–Teller (BET) theory. The BET specific surface area was extrapolated from the linear region of the BET graph and the total pore volume was obtained from the data point at around P/P0 = 0.992.
Fig. 2A and B show electron microscopy images of as-synthesized HMCs. It is clear that HF etching removes the SiO2 core to yield a hollow structure that retains a spherical morphology with a diameter of 98 ± 20 nm and a shell thickness of 3.8 ± 0.4 nm (Fig. 2B). After addition of Mn(CH3CO2)2 and NaOH followed by 5 h of sonication, the morphology of the HMCs remained intact (Fig. 2C). The SEM image reveals a rough surface on the modified HMCs that presumably arises from the deposition of NPs. Higher magnification inspection using TEM (Fig. 2D) reveals Mn3O4 NPs on the surfaces and within the HMCs. The average NP size was determined to be 3.8 ± 0.5 nm (Fig. S6†). The crystallinity of the Mn3O4 NPs is clear from the selected area electron diffraction (SAED) pattern (Fig. 3A), which shows rings with d-spacings of 0.312, 0.282, 0.251, 0.204, 0.182, and 0.155 nm, that we confidently assigned to (112), (103), (211), (220), (105), and (224) planes, respectively, of hausmannite Mn3O4 (JCPDS card 24-0734).25,26 These d-spacings were also directly observed via HRTEM (Fig. 3B). PXRD analysis (Fig. 3C) further confirms the reflections are readily indexed to hausmannite Mn3O4.26
Fig. 2 Representative SEM secondary electron (SE) (A) and TEM bright field (BF) (B) images of HMC; SEM SE (C) and TEM BF (D) images of Mn3O4@HMC. |
Fig. 3 Selected area electron diffraction pattern (A), HRTEM image (B), and PXRD pattern (C) for Mn3O4@HMC. |
EDX mapping of the HMCs (Fig. S1†) shows the N signal overlaps with the bright field image, which is consistent with N being incorporated into the carbon matrix. The oxygen present in the carbon matrix can be reasonably attributed to residual oxygen remaining after the pyrolysis of polydopamine. Fig. 4 shows EDX mapping for the Mn3O4@HMC. The C, Mn, and O signals all overlap with the STEM HAADF image. EDX analysis shows that the NPs are uniformly distributed and have a composition consistent with Mn3O4.
The Mn3O4@HMC hybrid was also characterized using XP and FTIR spectroscopy. As expected, the XP survey spectrum (Fig. S2†) shows emissions associated with Mn, C, N, and O. The deconvoluted N 1s XP spectrum (Fig. 5A) for HMC shows pyridinic and pyrrolic N at 398.2 eV and 400.5 eV, respectively. After incorporation of Mn3O4, the pyridinic N emission did not shift, while the pyrrolic N peak shifted from 400.5 eV to 399.9 eV. This observation can reasonably be attributed to pyrrolic N interacting with the Mn3O4 NPs and is the subject of ongoing investigation in our laboratory. In the deconvoluted C 1s spectrum (Fig. 5B), the CO emission shifted to a higher binding energy and merged with the O–CO feature, consistent with the CO species also interacting with the Mn3O4 surface through the oxygen. The deconvoluted Mn 2p spectrum (Fig. 5C) of the hybrids shows Mn 2p3/2 at 641.4 eV, confidently assigned to Mn3O4.27 A satellite peak corresponding to Mn3O4 was also observed at 649.2 eV. The Mn 3s spectrum (Fig. 5D) shows a signature 5.4 eV splitting of the 3s emission that is commonly attributed to the presence of Mn3+.27–29
FTIR spectra for HMC and Mn3O4@HMC are shown in Fig. 6. The HMC spectrum shows an absorption at 3434 cm−1 that is attributed to surface O–H and N–H stretching.30 The feature at 2925 cm−1 arises from C–H stretching.30 Features at 1585 and 1435 cm−1 correspond to C–N bending and heterocyclic stretching, respectively, and the feature at 1172 cm−1 is assigned to heterocyclic N–H in-plane deformation breathing.30 In addition to all the HMC spectral features, the Mn3O4@HMC spectrum shows an absorption at 524 cm−1 corresponding to Mn–O bonding.31 The C–N bending and heterocyclic stretching features in the HMC spectrum are qualitatively sharper and located at lower energy (i.e., 1556 and 1410 cm−1, respectively) for the hybrid, which is consistent with HMC-NP interactions suggested by the N-region of the XP spectra (Table 1).
Catalyst | BET surface area (m2 g−1) | Pore diameter (nm) | Pore volume (cm3 g−1) |
---|---|---|---|
Pristine HMC | 498 | 29 | 3.33 |
Mn3O4@HMC | 201 | 24 | 1.28 |
The textural properties of the pristine HMC and Mn3O4@HMC were evaluated by nitrogen sorption analysis. BET surface area, pore diameter, and pore volume were acquired. Isotherms (Fig. S3 and S4†) show a distinct hysteresis loop was observed at high relative pressure for the pristine HMCs giving an average pore diameter is 29 nm, consistent with a mesoporous material.22 In addition, they exhibit a high BET surface area of 498 cm2 g−1 was determined. After incorporation of Mn3O4 NPs, the BET surface area and pore volume drop to 201 cm2 g−1 and 1.28 cm3 g−1, respectively. This observation can be attributed to Mn3O4 NPs occupying the space inside the HMCs.
The ORR activity of Mn3O4@HMC was first assessed using cyclic voltammetry (CV) in both Ar- and O2-saturated 0.1 M KOH aqueous solution using a rotating disk electrode (Fig. 7A). For these measurements, a 0.1 M KOH aqueous electrolyte was used to maximize O2 solubility and minimize background current.32 The Ar-saturated cyclic voltammogram is featureless and only shows capacitive current; in contrast, the cyclic voltammogram from the O2-saturated system shows a cathodic current at −0.17 V. This observation indicates that the ORR reaction occurs on the Mn3O4@HMC. The ORR performance was subsequently evaluated using standard linear sweep voltammetry (LSV) in 1 M KOH aqueous electrolyte. Electrodes were prepared by impregnating a circular GDL with pristine HMC or Mn3O4@HMC. LSV curves of pristine HMC, Mn3O4 NPs, Mn3O4@HMCs, and Pt–Ru are shown in Fig. 7B. The onset potential (defined as the potential at which the current density reaches 10 mA cm−2) for the hybrid material was markedly improved from −0.156 V for the pristine HMC to −0.082 V; this suggests synergistic effects resulting from the combination of HMCs with Mn3O4 NPs. This performance is comparable to that of Pt–Ru (−0.077 V). In addition, the maximum current density obtained for Mn3O4@HMC is 198.1 mA cm−2 and exceeds that of pristine HMC (i.e., 167.6 mA cm−2); it is marginally better than that of the Pt–Ru (i.e., 196.5 mA cm−2). Based upon these observations, Mn3O4@HMC exhibits improved catalytic activity relative to pristine HMC and meets or exceeds the performance of our Pt–Ru benchmark.
To investigate the kinetics and catalytic mechanism, LSV was also performed using a Mn3O4@HMC-coated rotating disk electrode (RDE) at a scan rate of 5 mV s−1 and predefined rotation rates (i.e., 225, 400, 625, 900, 1225, and 1600 rpm) in O2-saturated 0.1 M KOH electrolyte (Fig. 8A). In Fig. 8A, kinetically controlled (0 to −0.1 V), kinetic-diffusion controlled (−0.1 to −0.3 V), and diffusion controlled (−0.3 to −0.7 V) regions were observed. The Koutecky–Levich (K–L) eqn (1) was used to determine the number of electrons transferred:
(1) |
ik = nFAkCO2 | (2) |
B = 0.62nAFCO2ν−1/6D02/3 | (3) |
O2 + 2H2O + 2e− → HO−2 + OH− | (4) |
2HO−2 → 2OH− + O2 | (5) |
O2 + 2H2O + 4e− → 4OH− | (6) |
Fig. 8 (A) ORR LSV curves for Mn3O4@HMC in O2-saturated 0.1 M KOH at scan rate of 5 mV s−1 using RDE; (B) K–L plots for Mn3O4@HMC extrapolated in the potential range of 0.4–0.6 V. |
Mn3O4@HMC and pristine HMC were incorporated into separate primary Zn–air batteries as the air electrode catalyst to evaluate their performance. Discharge rate tests were performed at predefined current densities (2, 5, 10, and 20 mA cm−2). In all cases, the rate discharge curves (Fig. 9A) show the discharge potential of the Mn3O4@HMC was markedly better compared with pristine HMC. Mn3O4@HMC also outperformed Pt–Ru at the tested current densities, indicating its superior ORR catalytic activity. At 10 and 20 mA cm−2, the discharge potentials were 1.26 and 1.22 V, respectively for Mn3O4@HMC. Our benchmark Pt–Ru catalyst only exhibited 1.25 and 1.20 V discharge potential at 10 and 20 mA cm−2, respectively. Furthermore, the performance of the present Mn3O4@HMC outperforms many transition metal oxide and carbon nanomaterial hybrids (Table S1†). This is also the first time N-doped hollow carbon nanospheres have been combined with transition metal oxide nanoparticles as a Zn–air battery catalyst. Conventional transition metal nanoparticle synthesis often involves high temperature annealing (>300 °C), which is likely to destroy the delicate feature of the HMC (average thickness = 3.8 nm).36–38 By implementing the sonication procedure, the potential damage to the HMC active sites was avoided.
Fig. 9 (A) Rate discharge curves; (B) polarization and power density curves for a primary Zn–air battery using Mn3O4@HMC, pristine HMC, and Pt–Ru as ORR catalysts. |
Polarization and power density curves are shown in Fig. 9B. They highlight that the maximum power density was improved from 158 mW cm−2 to 183 mW cm−2 after the incorporation of Mn3O4 NPs. The maximum power density is much higher than the value obtained for our Pt–Ru benchmark (158 mW cm−2) under the same conditions. The hybrid catalyst also exhibits a much lower charge-transfer resistance, evidenced by the smaller size of the semi-circular region of EIS Nyquist plot (Fig. S8†). The excellent catalytic performance can be attributed to the exterior and interior surfaces of the HMCs being decorated with Mn3O4 NPs. The number of active sites was maximized and any synergistic effects were amplified.
The stability of the as-synthesized hybrid material was further investigated by cycling a rechargeable Zn–air battery at 20 mA cm−2 using Mn3O4@HMC on GDL as the ORR electrode. Each cycle was 30 min and 235 cycles (117.5 h) were performed. Ni foam was used as the OER electrode due to its high surface area, good OER catalytic activity, and stability.39 A tri-electrode configuration was utilized so that the ORR performance could be evaluated independently of the OER performance. As shown in the discharge/charge curves (Fig. 10A), the HMC hybrid initially exhibits a discharge potential of 1.21 V. After 117.5 h of cycling (235 cycles), the discharge potential decreased slightly to 1.17 V corresponding to a 3.3% loss. A similar tri-electrode battery, using Pt–Ru on GDL as the ORR electrode and Ni foam as OER electrode, was tested under the same conditions (Fig. 10B). The discharge potential for Pt–Ru decreased from 1.19 to 1.14 V, after 117.5 h of cycling, corresponding to a 4.2% change. This shows the comparable long-term durability of the Mn3O4@HMC compared with Pt–Ru. The morphology of the HMC hybrid was not evaluated after cycling due to the difficulty in separating the hybrid from the GDL.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0na00428f |
This journal is © The Royal Society of Chemistry 2020 |