DOI:
10.1039/D4DT01733A
(Paper)
Dalton Trans., 2024,
53, 14801-14810
Ni-doping optimized d-band center in bifunctional Fe2O3 modified by bamboo-like NCNTs as a cathode material for Zn–air batteries†
Received
14th June 2024
, Accepted 6th August 2024
First published on 7th August 2024
Abstract
During the development of Zn–air batteries, designing an affordable, efficient and stable electrocatalyst for both oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) presents a great challenge. Fe2O3 exhibits ORR and OER activities, but when used as a cathode material in Zn–air batteries, its activity requires further improvement. To achieve this goal, Ni is doped into Fe2O3 hexagonal nanorods, derived from a metal–organic framework (MOF) precursor, and further modified by N-doped carbon nanotubes. In ORR, its half-wave potential achieves 0.946 and 0.716 V in alkaline and neutral electrolytes, respectively. In OER, it requires 388 mV to obtain 10 mA cm−2 in an alkaline electrolyte. As illustrated by theoretical calculation, Ni-doping raises the d-band center of Fe2O3, which enhances its adsorption towards relevant oxygen species in electrocatalysis. This improves its ORR and OER activities. Based on these merits, the Zn–air battery is assembled with an alkaline electrolyte. At 10 mA cm−2, its specific capacity and energy density reach 819.8 mA h g−1 and 960.1 W h kg−1, respectively. This battery remains stable after a long time of charge and discharge. In neutral electrolytes, its promising discharge performance is also well retained. This work develops an effective approach to improve ORR and OER activities of Fe2O3-based cathode materials in Zn–air batteries.
Introduction
Recently, driven by the global fossil energy and environmental pollution crisis, the exploration for clean, renewable, reliable energy conversion techniques and devices has become a significant challenge for researchers.1,2 As a new generation of energy conversion devices, Zn–air batteries have gained broad attention owing to their great advantages such as environmental friendliness, excellent safety, high capacity and energy density.3,4 For Zn–air batteries with alkaline electrolyte, discharge is achieved through ORR, which occurs at the cathode where O2 is reduced to OH−.5–7 On the contrary, during charging, OER takes place, converting OH− back to O2.8–10 However, for most ORR and OER electrocatalysts, their inherently weak kinetic processes are too sluggish to achieve high performance in Zn–air batteries.11,12 Pt/C and RuO2 are promising ORR and OER electrocatalysts; however, their popularization in Zn–air batteries is still seriously restricted by their limited reserves, expensive price and poor stability.13–16 Consequently, the exploration for stable and efficient bifunctional electrocatalysts with rich reserves and low price is pressing for further development of Zn–air battery.17,18
In this aspect, transition metal oxides have attracted great interest owing to their special advantages.19–21 First, transition metal oxides exhibit favorable corrosion resistance in alkaline electrolytes during the electrochemical process.22,23 Secondly, due to the coexistence of multiple valence states, intensive interactions are generated in transition metal oxides, which improves electrochemical activity.24–26 Considering the above-mentioned facts, cheap Fe2O3-based electrocatalysts appear to be ideal substitutes for Pt/C and RuO2 in ORR/OER.27–29 However, before entering into practical application as a cathode material for real Zn–air batteries, additional improvements are required for Fe2O3.30–33 To improve ORR/OER activities, optimization of the interaction between electrocatalyst and relevant active oxygen species (such as O2, OOH*, O*, OH* or OH−) is a feasible approach.34–36 To achieve this goal, band structure adjustment through hetero component (such as transition metallic elements, Mn, Co, Cu, etc. and nonmetallic elements N, S, P, etc.) doping is a feasible strategy, which has achieved great success.37–39
For Fe2O3, when employed as a bifunctional ORR and OER electrocatalyst in a Zn–air battery, poor conductivity is a conspicuous weakness, which weakens electrocatalytic activity greatly.40,41 To resolve this problem, the combination of Fe2O3 and carbon materials, especially N-doped carbon materials can improve conductivity greatly.42,43 This enhances ORR and OER activities simultaneously.44–46 In the obtained composite electrocatalyst, carbon material plays an important role, which serves as a conductive carrier and active component.47–50 Compared with carbon material in other forms, carbon nanotubes (CNTs), especially N-doped carbon nanotubes (NCNTs) seem more attractive because of their special characteristics, such as optimized pathway in charge transportation and short diffusion distance for relevant oxygen species in ORR/OER.51–58 Furthermore, NCNTs also possess superiority in electrocatalysis. Therefore, the combination of Fe2O3 and NCNTs can obtain electrocatalysts with ideal ORR and OER activities.
As motivated by the above points, with MIL-88A, a MOF as a precursor, a bifunctional electrocatalyst (Ni/Fe-Fe2O3@NCNTs) was synthesized. In Ni/Fe-Fe2O3@NCNTs, Ni/Fe-Fe2O3 hexagonal nanorod is modified using bamboo-shaped NCNTs derived from dicyandiamide (DCDA) (Scheme 1). It shows good ORR activity. In alkaline electrolyte, its half-wave potential (E1/2) achieves as high as 0.946 V with limiting current density (Jd) 5.35 mA cm−2. In neutral electrolytes, the E1/2 value still remains at 0.716 V. Its OER activity is also promising and the overpotential (η10) is 388 mV with a Tafel slope of 89.1 mV dec−1. Theoretical calculation demonstrates that Ni doping elevates the d-band center, which enhances adsorption towards relevant active oxygen species in ORR and OER. With Ni/Fe-Fe2O3@NCNTs as a cathode material, a Zn–air battery is constructed. In alkaline electrolyte, its peak density reaches 170.1 mW cm−2. At 10 mA cm−2, its specific capacity and energy density are 819.8 mA h g−1 and 960.1 W h kg−1, respectively. In neutral electrolytes, the discharge performance of the Zn–air battery is also well retained.
|
| Scheme 1 Schematic diagram of the synthetic procedure for Ni/Fe-Fe2O3@NCNTs. | |
Experimental section
Synthesis of Ni-doped MIL-88A
At first, FeCl3·6H2O (6 mmol, 1.62 g) and Ni(NO3)2·6H2O (0.5 mmol, 0.15 g) were dissolved in deionized water (30 mL). Then, fumaric acid (6 mmol, 0.7 g) was added. The obtained solution was stirred at room temperature for 30 min and transferred into a 50 mL stainless-steel autoclave. After heating at 70 °C for 16 h, the products were synthesized, which were washed with ethanol, and deionized water several times and dried in a vacuum oven at 80 °C for 12 h. To obtain neat MIL-88A, Ni(NO3)2·6H2O was not added.
Synthesis of Ni/Fe-Fe2O3@NCNTs
DCDA (1.5 g, on the gas inlet side) and Ni-doped MIL-88A (0.1 g, on the gas outlet side) were dispersed on both ends of the porcelain boat. After heating at 800 °C under N2 protection for 3 h at a rate of 5 °C min−1, black powder of Ni/Fe-Fe2O3@NCNTs was obtained. The products were washed using ethanol and deionized water several times and dried in a vacuum oven at 80 °C for 12 h. When MIL-88A was used instead of Ni-doped MIL-88A, Fe-Fe2O3@NCNTs was synthesized.
Results and discussion
Structure and morphology of Ni/Fe-Fe2O3@NCNTs
The structure of the calcined product was characterized by powder X-ray diffraction (PXRD). The diffraction peaks at 23.8, 30.2, 35.6, 43.3, 57.4 and 62.9° correspond with (210), (220), (311), (400), (511) and (440) planes of Fe2O3, respectively (JCPDS No. 39-1346) (Fig. 1a). The left two peaks at 44.6 and 64.9° originate from (110) and (200) planes of FeNi alloy (JCPDS No. 37-0474), respectively. Raman spectroscopy was also employed to study the structure. For Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs, two distinct peaks appear at 1350 and 1580 cm−1, which correspond with D and G bands, respectively, (Fig. 1b) were observed.59 In general, the D band reflects the content of carbon in the defect form, while the G band represents carbon in the graphic phase. The ratio between D and G bands (ID/IG) can be used to evaluate the degree of graphitization in carbon.60 For Ni/Fe-Fe2O3@NCNTs, the ID/IG value is 0.96, which demonstrates that NCNTs exist in a high degree of graphitization. This improves conductivity and ORR/OER activities.
|
| Fig. 1 (a) PXRD pattern of Ni/Fe-Fe2O3@NCNTs, standard FeNi alloy and Fe2O3; (b) Raman spectra of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs; (c) high-resolution Ni 2p XPS spectrum of Ni/Fe-Fe2O3@NCNTs; (d) high-resolution Fe 2p XPS spectra of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs; high-resolution XPS spectrum of Ni/Fe-Fe2O3@NCNTs, (e) N 1s and (f) O 1s. | |
The composition and valence states of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs were studied by X-ray photoelectron spectroscopy (XPS). For Ni/Fe-Fe2O3@NCNTs, C 1s, N 1s, O 1s, Fe 2p and Ni 2p peaks emerged in the survey spectrum (Fig. S1†). As for Fe-Fe2O3@NCNTs, Ni 2p peaks disappear completely. A high-resolution spectrum was used to analyze the valence states of the above five components. In the high-resolution Ni 2p spectrum, two doublet peaks emerge at 853.1, 870.6 and 855.5, 873.7 eV (Fig. 1c). They match well with Ni 2p3/2 and Ni 2p1/2 in Ni0 and Ni2+ states, respectively.61 The appearance of Ni2+ originates from the oxidation of FeNi alloy. The other two peaks at 861.8 and 880.2 eV can be ascribed to satellite peaks. In the high-resolution Fe 2p spectrum, the first pair of peaks located at 710.8 and 724.3 eV originate from Fe 2p3/2 and Fe 2p1/2, respectively, in the Fe2+ phase (Fig. 1d).62 The doublet peaks at 712.8 and 726.9 eV correspond with Fe 2p3/2 and Fe 2p1/2, respectively, of Fe3+. The other two are satellite peaks. Furthermore, compared with Fe-Fe2O3@NCNTs, Fe 2p peaks shift to a negative energy region, which demonstrates in Ni/Fe-Fe2O3@NCNTs, the electrons move from Ni to Fe2O3. In the high-resolution N 1s spectrum, the peaks at 398.8, 400.6 and 401.8 eV match with pyridinic N, pyrrolic N and graphitic N, respectively (Fig. 1e).63 High-resolution O 1s spectrum reveals that lattice oxygen (Oα) and oxygen vacancies (Oβ) coexist in Fe2O3 (Fig. 1f).64 In the high-resolution C 1s spectrum, three peaks appear at 284.6, 285.3 and 286.3 eV (Fig. S2†). They correspond to C–C, C–N and CN, respectively65
The morphology of Ni/Fe-Fe2O3@NCNTs was observed with scanning electron microscopy (SEM) at first. The SEM image illustrates that the Ni/Fe-Fe2O3@NCNTs hexagonal nanorod shows uniform size (Fig. S3†). The length of hexagonal nanorod ranges from 1.2 to 2 μm with a diameter of about 250 to 400 nm (Fig. 2a). It shows a similar appearance to MIL-88(A), besides surface-covered NCNTs derived from DCDA (Fig. S4†). The length of NCNTs is 120 to 200 nm with a diameter of 10 to 15 nm (Fig. 2b). Transmission electron microscopy (TEM) was also used to observe its morphology. TEM image reveals that hexagonal nanorod is composed of small Ni/Fe-Fe2O3 nanoparticles of about 20 to 40 nm size and NCNTs grow homogeneously on their surface (Fig. 2c). NCNTs exhibit a bamboo-shaped appearance with small particles residing in the channel (Fig. 2d). In NCNTs, Ni/Fe-Fe2O3 particles with the size about 10 to 15 nm reside homogeneously. With high-resolution transmission electron microscopy (HRTEM) images, more information is obtained. The lattice fringes of about 0.25 nm match with the (311) plane of Fe2O3 (Fig. 2e). Based on energy-dispersive X-ray fluorescence spectroscopy (EDX), C, N, O, Fe and Ni disperse evenly on Ni/Fe-Fe2O3@NCNTs (Fig. 2f). Furthermore, the EDX also confirms the content of Ni as 0.3%, which also proves that Ni was doped successfully.
|
| Fig. 2 (a) and (b) SEM images of Ni/Fe-Fe2O3@NCNTs; (c) and (d) TEM images of Ni/Fe-Fe2O3@NCNTs; (e) HRTEM image of Ni/Fe-Fe2O3@NCNTs; and (f) EDX of C, N, O, Fe and Ni in Ni/Fe-Fe2O3@NCNTs. | |
ORR activity study in alkaline electrolyte
The ORR activity was studied with the rotating disk electrode (RDE) technique. In the 0.1 M KOH electrolyte saturated by O2, Ni/Fe-Fe2O3@NCNTs exhibits more excellent ORR activity than Fe-Fe2O3@NCNTs and Pt/C. Based on linear sweep voltammetry (LSV) curves, its E1/2 achieves 0.946 V, which is higher than that of Fe-Fe2O3@NCNTs (0.909 V) and Pt/C (0.896 V) (Fig. 3a). The Jd value of Ni/Fe-Fe2O3@NCNTs is 5.35 mA cm−2 and this is larger than that of Pt/C (5.03 mA cm−2). Its merit over the other two competitors is also revealed by the Tafel slope. The slope of Ni/Fe-Fe2O3@NCNTs (68.79 mV dec−1) is lower than that of Fe-Fe2O3@NCNTs (83.68 mV dec−1) and Pt/C (76.8 mV dec−1), which implies that Ni/Fe-Fe2O3@NCNTs can obtain high current density at low potential (Fig. 3b). ORR activity was also evaluated by cyclic voltammetry (CV). Reductive peaks of Ni/Fe-Fe2O3@NCNTs, Fe-Fe2O3@NCNTs and Pt/C are located at 0.897, 0.801 and 0.811 V, respectively (Fig. 3c). This also implies that Ni/Fe-Fe2O3@NCNTs shows better ORR activity than Fe-Fe2O3@NCNTs and Pt/C. The ORR activity of Ni/Fe-Fe2O3@NCNTs is comparable with those of materials reported recently (Table S1†).
|
| Fig. 3 (a) LSV curves, (b) Tafel slopes, (c) CV curves of Ni/Fe-Fe2O3@NCNTs, Fe-Fe2O3@NCNTs and Pt/C; (d) current density as the function of scanning rates for Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs; (e) LSV curves before and after 5000 cycles of the ORR test of Ni/Fe-Fe2O3@NCNTs; (f) the chronoamperometric curve of Ni/Fe-Fe2O3@NCNTs for 12 h in ORR. | |
To further explore the ORR activity of different electrocatalysts, electrochemically active surface area (ECSA) is an important parameter, which can be reflected by a double-layer capacitor (Cdl). For Ni/Fe-Fe2O3@NCNTs, its Cdl value reaches 12.56 mF cm−2, which is higher obviously than Fe-Fe2O3@NCNTs (8.51 mF cm−2) (Fig. 3d and Fig. S5, S6†). The results indicate that Ni/Fe-Fe2O3@NCNTs can provide more active sites in ORR. In ORR, stability is also very important for Ni/Fe-Fe2O3@NCNTs. Its E1/2 only shifts negatively about 10 mV after 5000 cycles LSV tests (Fig. 3e). The stability of Ni/Fe-Fe2O3@NCNTs can also be judged by chronoamperometric curves at a constant potential. After a continuous ORR test for 12 h, the current density still remains at 88.59% (Fig. 3f). All these facts confirm its good stability in ORR.
In ORR, electron transfer number (n) is a significant parameter, which can be measured using the rotating disk electrode (RDE). For Ni/Fe-Fe2O3@NCNTs, when rotating speed increases from 400 to 2500 rpm, current density also increases (Fig. 4a). Simulated from the above LSV curves, the Koutecky–Levich (K–L) plots at different potentials were obtained (Fig. 4b). At 0.5, 0.6 and 0.7 V (vs. RHE), its electron transfer numbers are 4.04, 4 and 3.92, respectively (Fig. 4c). This implies that Ni/Fe-Fe2O3@NCNTs shows 4e-character in ORR. Based on a similar method, the electron transfer numbers of Fe-Fe2O3@NCNTs are also calculated, which are 4.06, 3.96 and 3.94 at the above potentials (Fig. S7, S8 and S9,† respectively). This implies that Ni-doping does not affect electron transfer numbers. The ORR activities of Ni/Fe-Fe2O3@NCNTs and Fe2O3@NCNTs were further studied with a rotating ring disk electrode (RRDE). The disk current density of Ni/Fe-Fe2O3@NCNTs is 5.46 mA cm−2, which is higher than that of Fe-Fe2O3@NCNTs (Fig. 4d and Fig. S10†). This further implies the superior ORR activity of Ni/Fe-Fe2O3@NCNTs over Fe-Fe2O3@NCNTs after Ni-doping. H2O2 selectivity of Ni/Fe-Fe2O3@NCNTs fluctuates between 1.93 and 11.2%, while the electron transfer number ranges from 3.78 to 3.96 (Fig. 4e and f). As for Fe-Fe2O3@NCNTs, it also shows low H2O2 selectivity from 1.31 to 5.75% with the electron transfer number increasing from 3.88 to 3.97 (Fig. S11 and S12†). Both RDE and RRDE results confirm the 4e-character of Ni/Fe-Fe2O3@NCNTs.
|
| Fig. 4 (a) LSV curves at different scanning rates of Ni/Fe-Fe2O3@NCNTs; (b) K–L equation simulation for Ni/Fe-Fe2O3@NCNTs; (c) n values of Ni/Fe-Fe2O3@NCNTs under different potentials; (d) RRDE voltammograms of Ni/Fe-Fe2O3@NCNTs; (e) n value and (f) H2O2 selectivity of Ni/Fe-Fe2O3@NCNTs. | |
ORR activity study in neutral electrolyte
The ORR activity of Ni/Fe-Fe2O3@NCNTs was also studied in neutral electrolytes. As the CV curve illustrates it exhibits a distinct oxygen reduction peak in 0.1 M PBS (pH = 7.4) saturated by O2 (Fig. S13†). The peak is located at 0.736 V (vs. RHE), which is obviously higher than that of Pt/C. This implies that Ni/Fe-Fe2O3@NCNTs shows more excellent ORR activity than Fe-Fe2O3@NCNTs under neutral conditions. In ORR, the advantage of Ni/Fe-Fe2O3@NCNTs over Fe-Fe2O3@NCNTs can be revealed by the LSV curve. The E1/2 and Jd values of Ni/Fe-Fe2O3@NCNTs reach 0.716 V and 4.47 mA cm−2, (Fig. 5a); As for Pt/C, the corresponding values are 0.616 V and 4.4 mA cm−2, respectively. For OR electrocatalysts, the stability is also very important in practical applications. For Ni/Fe-Fe2O3@NCNTs, the E1/2 value merely shifts negatively about 20 mV before and after 5000 cycles of the LSV tests (Fig. 5b). This illustrates that its excellent stability is well retained in neutral electrolytes.
|
| Fig. 5 (a) LSV curves of Ni/Fe-Fe2O3@NCNTs and Pt/C in 0.1 M PBS; (b) LSV curves before and after 5000 cycles tests of Ni/Fe-Fe2O3@NCNTs in 0.1 M PBS; (c) LSV curves at different scanning rates of Ni/Fe-Fe2O3@NCNTs in 0.1 M PBS; (d) n values of Ni/Fe-Fe2O3@NCNTs under different potentials in 0.1 M PBS; (e) RRDE voltammograms and (f) n value of Ni/Fe-Fe2O3@NCNTs in 0.1 M PBS. | |
The electron transfer number of Ni/Fe-Fe2O3@NCNTs was simulated based on the K–L equation with the RDE technique. The rotating speeds were 400, 625, 900, 1225, 1600, 2025 and 2500 rpm in the LSV tests (Fig. 5c). With increasing rotating speed, current density also increases. Simulated from the K–L equation, the linear relationship between J−1 and ω−1/2 is satisfactory (Fig. S14†). At 0.1, 0.2 and 0.3 V (vs. RHE), the electron transfer numbers are 3.86, 3.83 and 3.74, respectively (Fig. 5d). The RRDE test was also conducted to further evaluate the electron transfer number (Fig. 5e). The electron transfer number of Ni/Fe-Fe2O3@NCNTs ranges from 3.78 to 3.96 and H2O2 selectivity is in the scope of 1.93 and 11.2% (Fig. 5f and Fig. S15†). Both RDE and RRDE results imply that 4e−-ORR dominates the oxygen reduction process in neutral electrolytes. The stability of Ni/Fe-Fe2O3@NCNTs was studied using the chronoamperometry method. After the 12 h test, the current density still keeps as high as 90.45% (Fig. S16†). This confirms the ideal ORR stability of Ni/Fe-Fe2O3@NCNTs in neutral electrolytes.
OER activity study in alkaline electrolyte
In general, the performance of the Zn–air battery is also influenced greatly by the OER activity of the electrocatalyst. For Ni/Fe-Fe2O3@NCNTs, its OER activity was studied in 1 M KOH. At 10 mA cm−2, its η10 is only 388 mV, which is higher slightly than RuO2 (Fig. 6a). As for Fe-Fe2O3@NCNTs, to obtain 10 mA cm−2, it overpotential achieves 483 mV, which is higher than Ni/Fe-Fe2O3@NCNTs. This confirms that Ni-doping can improve OER activity effectively. For an electrocatalyst, the advantage in OER can also be illustrated by the Tafel slope. As for Ni/Fe-Fe2O3@NCNTs, its Tafel slope is 89.1 mV dec−1 (Fig. 6b). This value is comparable to that of RuO2 (83.9 mV dec−1) and lower than that of Fe-Fe2O3@NCNTs (191.72 mV dec−1). The stability in OER was also studied in 1 M KOH. The LSV curves before and after the 2000 cycles LSV test almost coincide, which indicates its promising stability (Fig. 6c). The OER activity is comparable with recently reported materials (Table S2†). To evaluate the applicability of bifunctional ORR/OER electrocatalyst, the gap (Δ) between E1/2 and η10 is a vital index. For Ni/Fe-Fe2O3@NCNTs, Δ value is 0.67 V (Fig. 6d). This demonstrates that it is an excellent bifunctional electrocatalyst for Zn–air batteries.
|
| Fig. 6 (a) LSV curves of Ni/Fe-Fe2O3@NCNTs, Fe-Fe2O3@NCNTs and RuO2 in OER; (b) Tafel slopes of Ni/Fe-Fe2O3@NCNTs, Fe-Fe2O3@NCNTs and RuO2 in OER; (c) LSV curves of Ni/Fe-Fe2O3@NCNTs before and after 5000 cycle tests; (d) ΔE values for Ni/Fe-Fe2O3@NCNTs, Fe-Fe2O3@NCNTs and Pt/C + RuO2. | |
Mechanism study of ORR and OER
To analyze different electrocatalytic activities between Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs, density functional theory (DFT) calculation was employed. As for ORR, O2 is adsorbed on the electrocatalyst, which is reduced to OOH*, O*, OH* and OH− in turn (Table S3†).66 For Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs, the free energy of each reductive step is thermodynamically favorable downhill as U = 0 V (Fig. 7a and b). After U increases to 1.23 V, the first and second steps (O2 + H2O + e → OOH* + OH− and OOH* + e → O* + OH−) become uphill. It can be observed clearly that Ni/Fe-Fe2O3@NCNTs need lower energy than Fe-Fe2O3@NCNTs to complete the barrier. To evaluate ORR activity in theory, limiting potential (the highest potential to make all elementary reactions exothermic) is an important parameter, which is abbreviated as UL(ORR). For Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs, the UL(ORR) are 0.95 and 0.74 V. Higher UL(ORR) of Ni/Fe-Fe2O3@NCNTs implies that it exhibits more excellent ORR activity than Fe-Fe2O3@NCNTs. In OER, as U = 0 V, the free energy of each step is uphill for Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs (Fig. 7c and d). At U = 1.23 V, the first two reactions (OH− − e = OH*; OH* + OH− − e = H2O + O*) are still uphill and become the rate-determining steps. It is obvious that Ni/Fe-Fe2O3@NCNTs needs low energy to overcome the barrier. Furthermore, the UL(OER) (limiting potential in OER) of Ni/Fe-Fe2O3@NCNTs is 1.51 V, which is lower than that of Fe-Fe2O3@NCNTs with UL(OER) 1.78 V. This confirms superior OER activity of Ni/Fe-Fe2O3@NCNTs over Fe-Fe2O3@NCNTs.
|
| Fig. 7 (a) Free energy diagram in ORR for Ni/Fe-Fe2O3@NCNTs at 0, 0.95 and 1.23 V; (b) free energy diagram in ORR for Fe-Fe2O3@NCNTs at 0, 0.74 and 1.23 V; (c) free energy diagram in OER for Ni/Fe-Fe2O3@NCNTs at 0, 1.51 and 1.23 V; (d) free energy diagram in OER for Fe-Fe2O3@NCNTs at 0, 1.78 and 1.23 V; (e) DOS diagram of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs showing d-band center position; (f) valence band spectra of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs. | |
In both ORR and OER, the rate-determining steps of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs are the generation and reduction of OOH* and OH*. To accelerate these steps, intensive adsorption towards relevant oxygen species (O2, OOH*, OH− and OH*) becomes very significant. According to the d-band center theory, the adsorption towards oxygen species is determined by the position of the d-band center. If the d-band center approaches the Fermi energy level, adsorption is intensive. On the contrary, the adsorption becomes weak.67 For Fe-Fe2O3@NCNTs, the d-band center is located at −2.69 eV based on the density of states (DOS). The d-band center of Ni/Fe-Fe2O3@NCNTs shifts to −2.16 eV (Fig. 7e). The position of the valence band can also be employed to measure the d-band center.68 For Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs, the valence band is located at 1.31 and 1.57 eV, respectively, based on ultraviolet photoelectron spectrometer (UPS) (Fig. 7f). These imply that the d-band center of Ni/Fe-Fe2O3@NCNTs is closer to the Fermi energy level than that of Fe-Fe2O3@NCNTs, which reveals that Ni-doping intensifies adsorption towards oxygen species and improves ORR and OER activities.
The performance of a Zn–air battery
Due to excellent ORR and OER activities in an alkaline environment, a rechargeable Zn–air battery was constructed with Ni/Fe-Fe2O3@NCNTs as the cathode material. In this Zn–air battery, the Zn slice serves as an anode, while a mixed solution of KOH (6 M) and Zn(OAc)2 (0.2 M) acting as an electrolyte. The open-circuit voltage of this battery reaches 1.46 V (Fig. 8a and b). This is obviously higher than that obtained for the Zn–air batteries employing Fe-Fe2O3@NCNTs and Pt/C + RuO2 as cathode materials. Furthermore, the peak power density of this Zn–air battery achieves 170.1 mW cm−2, which is superior to Zn–air batteries built by Fe-Fe2O3@NCNTs (137.48 mW cm−2) and Pt/C + RuO2 (104.26 mW cm−2) (Fig. 8c). After linking two Zn–air batteries with series mode, green LED lights with rated voltage of about 2.5 V can be lit up and work continuously for more than 12 h (Fig. 8d). This battery also shows small gap between charge and discharge voltages (Fig. 8e). At 50 mA cm−2, the voltage gap is only 1.12 V. This is comparable with most of recently reported Zn–air batteries.
|
| Fig. 8 (a) Photograph of one Zn–air battery (with alkaline electrolyte) with an open circuit voltage of 1.46 V; (b) polar curves of Zn–air battery; (c) power density curves of Zn–air battery; (d) photograph of LED lights powered by two Zn–air batteries linked in series; (e) charge/discharge polar curves of Zn–air battery; (f) galvanostatic discharge curves at 1, 5 and 10 mA cm−2; (g) galvanostatic discharge curve of at 1, 2, 4, 8 and 10 mA cm−2; (h) recycled charge/discharge test at 1 mA cm−2 for 80 h. | |
The discharge performance of this Zn–air battery was studied with galvanostatic technique at 1, 5 and 10 mA cm−2 (Fig. 8f). At the above current densities, the specific capacities are 863.9, 838.5, 819.8 mA h g−1 and energy densities are 1105.8, 1006.2, 960.1 W h kg−1, respectively. These values are adjacent to the theoretical parameters of the Zn–air battery. In addition, the specific capacities of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs were studied at 5 mA cm−2. The specific capacity of Ni/Fe-Fe2O3@NCNTs is obviously higher than that of Fe-Fe2O3@NCNTs (755.9 mA h g−1) (Fig. S17†). Furthermore, the discharge performance of this Zn–air battery can be comparable with recently reported Zn–air batteries (Table S4†). For Zn–air battery, rate ability is also very important and this can be measured by galvanostatic technique at various current densities. Here, current densities were set at 1, 2, 4, 8 and 10 mA cm−2 (Fig. 8g). Discharge voltage attenuates gradually with increasing current densities and the voltage plateaus are located at 1.32, 1.28, 1.24, 1.18 and 1.14 V. After 5 h of the discharge experiment, the current density returns to 1 mA cm−2 and the voltage becomes 1.31 V with insignificant attenuation. This confirms the promising rate capacity of this Zn–air battery. The stability of the Zn–air battery was evaluated by long-time charge/discharge tests at 1 mA cm−2 (Fig. 8h). After continuous tests lasting 80 h, the charge/discharge voltages still remained constant. This illustrates its ideal stability.
Due to excellent ORR activity in the neutral electrolyte, a new Zn–air battery was built up with Ni/Fe-Fe2O3@NCNTs as the cathode material, while a mixed solution of NH4Cl (4 M) and KCl (1 M) serves as the electrolyte. For this new Zn–air battery, its open-circuit voltage reaches 1.45 V (Fig. 9a and b). When discharged at a similar voltage, it can provide a higher current density than a battery constructed by Pt/C and RuO2. The peak power density of this Zn–air battery is 95.7 mW cm−2 (Fig. 9c). This is obviously higher than the battery built by Pt/C and RuO2 (41.9 mW cm−2). Two Zn–air batteries linked with series mode can also light up green LED lights with a rated voltage of about 2.5 V (Fig. 9d). Rate capacity of this battery was studied under different current densities. After the discharge at 0.5, 1, 2, 4, 8 and 10 mA cm−2 for 6 h, the voltage can almost return to the original value (Fig. 9e). Discharge performance of this Zn–air battery was evaluated with galvanostatic technique at 5 and 10 mA cm−2 (Fig. 9f). The specific capacities are 812.9 and 802.4 mA h g−1 with energy density 646.3 and 579.4 W h kg−1, respectively. The above facts imply that Ni/Fe-Fe2O3@NCNTs has excellent ORR and OER activities and is a qualified cathode material in both alkaline and neutral electrolytes.
|
| Fig. 9 (a) Photograph of one Zn–air battery (using a neutral electrolyte) with an open circuit voltage of 1.45 V; (b) polar curves of Zn–air battery; (c) power density curves of Zn–air battery; (d) photograph of LED lights powered by two Zn–air batteries linked in series; (e) galvanostatic discharge curve at 0.5, 1, 2, 4, 8 and 10 mA cm−2; (f) galvanostatic discharge curves at 5 and 10 mA cm−2. | |
Conclusions
In summary, Ni/Fe-Fe2O3@NCNTs, a bifunctional electrocatalyst was synthesized with Fe-MOF as a precursor. In Ni/Fe-Fe2O3@NCNTs, NCNTs grow on Ni/Fe-doped Fe2O3 hexagonal nanorod homogenously. In ORR, Ni/Fe-Fe2O3@NCNTs exhibit high E1/2 and Jd in both alkaline and neutral electrolytes. In OER, it shows a low overpotential and Tafel slope. Theoretical calculation implies that Ni-doping promotes the d-band center in Fe2O3, which intensifies its interaction with relevant oxygen species during electrocatalysis. Using Ni/Fe-Fe2O3@NCNTs as a cathode material, the Zn–air battery is assembled in an alkaline electrolyte. It shows promising charge and discharge performance with a high specific capacity, energy density and power density. This battery also exhibits promising stability during charge and discharge. In neutral electrolytes, the discharge performance of the Zn–air battery is also well retained. We anticipate band structure adjustment by hetero component doping will be popularized to improve ORR and OER activities for Zn–air batteries.
Data availability
The data in this manuscript will be available on request.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (22171039); and Fundamental Research Funds for the Central University (N2209005).
References
- Y. Zhao, Z. Zhang, L. Liu, Y. Wang, T. Wu, W. Qin, S. Liu, B. Jia, H. Wu, D. Zhang, X. Qu, G. Qi, E. P. Giannelis, M. Qin and S. Guo, S and O Co-coordinated Mo single sites in hierarchically porous tubes from sulfur-enamine copolymerization for oxygen reduction and evolution, J. Am. Chem. Soc., 2022, 144(45), 20571–20581 CrossRef CAS PubMed .
- D. Takimoto, S. Toma, Y. Suda, T. Shirokura, Y. Tokura, K. Fukuda, M. Matsumoto, H. Imai and W. Sugimoto, Platinum nanosheets synthesized via topotactic reduction of single-layer platinum oxide nanosheets for electrocatalysis, Nat. Commun., 2023, 14(1), 14–19 CrossRef PubMed .
- Q. Lu, J. Yu, X. Zou, K. Liao, P. Tan, W. Zhou, M. Ni and Z. Shao, Self–catalyzed growth of Co, N–codoped CNTs on carbon–encased CoSx surface: A noble–metal–free bifunctional oxygen electrocatalyst for flexible solid Zn-air batteries, Adv. Funct. Mater., 2019, 29(38), 1904481 CrossRef .
- F. Shahbazi Farahani, M. S. Rahmanifar, A. Noori, M. F. El-Kady, N. Hassani, M. Neek-Amal, R. B. Kaner and M. F. Mousavi, Trilayer metal-organic frameworks as multifunctional electrocatalysts for energy conversion and storage applications, J. Am. Chem. Soc., 2022, 144(8), 3411–3428 CrossRef CAS PubMed .
- M. Wu, G. Zhang, H. Tong, X. Liu, L. Du, N. Chen, J. Wang, T. Sun, T. Regier and S. Sun, Cobalt(II) oxide nanosheets with rich oxygen vacancies as highly efficient bifunctional catalysts for ultra-stable rechargeable Zn-air flow battery, Nano Energy, 2021, 79, 105409 CrossRef CAS .
- Y. Zhou, R. Lu, X. Tao, Z. Qiu, G. Chen, J. Yang, Y. Zhao, X. Feng and K. Müllen, Boosting oxygen electrocatalytic activity of Fe-N-C catalysts by phosphorus incorporation, J. Am. Chem. Soc., 2023, 145(6), 3647–3655 CrossRef CAS PubMed .
- S. Chen, L. Ma, S. Wu, S. Wang, Z. Li, A. A. Emmanuel, M. R. Huqe, C. Zhi and J. A. Zapien, Uniform virus-like Co-N-Cs electrocatalyst derived from prussian blue analog for stretchable fiber-shaped Zn-air batteries, Adv. Funct. Mater., 2020, 30(10), 1908945 CrossRef CAS .
- W.-J. Niu, Q.-Q. Sun, J.-Z. He, J.-L. Chen, B. Gu, M.-J. Liu, C.-C. Chung, Y. Wu and Y.-L. Chueh, Zeolitic imidazolate framework-derived copper single atom anchored on nitrogen-doped porous carbon as a highly efficient electrocatalyst for the oxygen reduction reaction toward Zn-air battery, Chem. Mater., 2022, 34(9), 4104–4114 CrossRef CAS .
- J. Hao, X. Li, S. Zhang, F. Yang, X. Zeng, S. Zhang, G. Bo, C. Wang and Z. Guo, Designing dendrite–free zinc anodes for advanced aqueous zinc batteries, Adv. Funct. Mater., 2020, 30(30), 2001263 CrossRef CAS .
- Z. Song, J. Ding, B. Liu, X. Liu, X. Han, Y. Deng, W. Hu and C. Zhong, A rechargeable Zn-air battery with high energy efficiency and long life enabled by a highly water–retentive gel electrolyte with reaction modifier, Adv. Mater., 2020, 32(22), 1908127 CrossRef CAS PubMed .
- M. Liu, Z. Liu, W. Chen, Z. Liu, Z. Li, X. Pi, Q. Du, X. Lai, Y. Xia and Y. Li, NiFe nanoparticle-encapsulated ultrahigh-oxygen-doped carbon layers as bifunctional electrocatalysts for rechargeable Zn-air batteries, Inorg. Chem., 2023, 62(28), 11199–11206 CrossRef CAS PubMed .
- J. Ran, T. Wang, J. Zhang, Y. Liu, C. Xu, S. Xi and D. Gao, Modulation of electronics of oxide perovskites by sulfur doping for electrocatalysis in rechargeable Zn-air batteries, Chem. Mater., 2020, 32(8), 3439–3446 CrossRef CAS .
- M. Yang, K. Ge, Y. Wu, B. Zhang and J. Duan, Synergistic catalysis of cobalt tetroxide and bamboo-shaped carbon nanotubes doped with nitrogen for oxygen reduction in Zn-air batteries, Inorg. Chem., 2023, 62(33), 13378–13386 CrossRef CAS PubMed .
- Q. Xu, L. Peng, K. Luo, J. Zhong, C. Zhang and D. Yuan, CoSe2 nanoparticles anchored on CoNC carbon nanoplates as bifunctional electrocatalyst for flexible rechargeable Zn-air batteries, J. Colloid Interface Sci., 2023, 643, 73–81 CrossRef CAS PubMed .
- C. Zhou, X. Chen, S. Liu, Y. Han, H. Meng, Q. Jiang, S. Zhao, F. Wei, J. Sun, T. Tan and R. Zhang, Superdurable bifunctional oxygen electrocatalyst for high-performance zinc-air batteries, J. Am. Chem. Soc., 2022, 144(6), 2694–2704 CrossRef CAS PubMed .
- S. Xie, H. Jin, C. Wang, H. Xie, Y.-R. Lu, T.-S. Chan, W. Yan, S. Jin and H. Ji, A comparison study on single metal atoms (Fe, Co, Ni) within nitrogen-doped graphene for oxygen electrocatalysis and rechargeable Zn-air batteries, Chin. Chem. Lett., 2023, 34(6), 107681 CrossRef CAS .
- E. Y. Choi, D. E. Kim, S. Y. Lee, C. B. Park and C. K. Kim, Cobalt nanoparticles-encapsulated holey nitrogen-doped carbon nanotubes for stable and efficient oxygen reduction and evolution reactions in rechargeable Zn-air batteries, Appl. Catal., B, 2023, 325, 122386 CrossRef CAS .
- J. Nie, M. Dong, G. Chen, N. Wang, J. Nie and G. Ma, Biomass-based hierarchical porous ORR and OER bifunctional catalysts with strong stability for Zn-air batteries, ACS Sustain, Chem. Eng., 2023, 11(30), 11161–11171 CAS .
- D. Meng, C. Zhang, Y. Liang, W. Qiu, F. Kong, X. He, M. Chen, P. Liang and Z. Zhang, Electrospun cobalt prussian blue analogue-derived nanofibers for oxygen reduction reaction and lithium-ion batteries, J. Colloid Interface Sci., 2021, 599, 280–290 CrossRef CAS PubMed .
- Y. Yang, R. Zeng, H. Paik, D.-Y. Kuo, D. G. Schlom, F. J. DiSalvo, D. A. Muller, J. Suntivich and H. D. Abruña, Epitaxial thin-film spinel oxides as oxygen reduction electrocatalysts in alkaline media, Chem. Mater., 2021, 33(11), 4006–4013 CrossRef CAS .
- E. L. Ribeiro, E. M. Davis, M. Mokhtarnejad, S. Hu, D. Mukherjee and B. Khomami, MOF-derived PtCo/Co3O4 nanocomposites in carbonaceous matrices as high-performance ORR electrocatalysts synthesized via laser ablation techniques, Catal. Sci. Technol., 2021, 11(9), 3002–3013 RSC .
- K. Min, S. Kim, E. Lee, G. Yoo, H. C. Ham, S. E. Shim, D. Lim and S.-H. Baeck, A hierarchical Co3O4/CoS microbox heterostructure as a highly efficient bifunctional electrocatalyst for rechargeable Zn-air batteries, J. Mater. Chem. A, 2021, 9(32), 17344–17352 RSC .
- D. Chen, W. Cao, J. Liu, J. Wang, X. Li and L. Jiang, Filling the in situ-generated vacancies with metal cations captured by C-N bonds of defect-rich 3D carbon nanosheet for bifunctional oxygen electrocatalysis, J. Energy Chem., 2021, 59, 47–54 CrossRef CAS .
- X. Zhuang, Y. Zhou, Z. Jiang, X. Yao and X.-Y. Yu, Synergetic electronic modulation and nanostructure engineering of heterostructured RuO2/Co3O4 as advanced bifunctional electrocatalyst for zinc-air batteries, J. Mater. Chem. A, 2021, 9(47), 26669–26675 RSC .
- Y. Li, J. Li, Y. Dai, X. Li, C. Shao, Y. Sun, R. Wang and J. Zou, Co3O4@carbon with high Co2+/Co3+ ratios derived from ZIF-67 supported on N-doped carbon nanospheres as stable bifunctional oxygen catalysts, Mater, Today Energy, 2021, 21, 100737 CAS .
- S. Das, S. Bhattacharjee, S. Mondal, S. Dutta, N. Bothra, S. K. Pati and S. Bhattacharyya, Bimetallic zero-valent alloy with measured high-valent surface states to reinforce the bifunctional activity in rechargeable zinc-air batteries, ACS Sustain. Chem. Eng., 2021, 9(44), 14868–14880 CrossRef CAS .
- L. Bo, F. Nian, L. Pu, P. Li, Y. Hu and J. Tong, Simply embedding α-Fe2O3/Fe3O4 nanoparticles in N-doped graphitic carbon polyhedron layered arrays as excellent electrocatalyst for rechargeable Zn-air battery, Sep. Purif. Technol., 2023, 312, 123413 CrossRef CAS .
- F. Xiao, X. Liu, C.-J. Sun, I. Hwang, Q. Wang, Z. Xu, Y. Wang, S. Zhu, H.-W. Wu, Z. Wei, L. Zheng, D. Cheng, M. Gu, G.-L. Xu, K. Amine and M. Shao, Solid-state synthesis of highly dispersed nitrogen-coordinated single iron atom electrocatalysts for proton exchange membrane fuel cells, Nano Lett., 2021, 21(8), 3633–3639 CrossRef CAS PubMed .
- K. Maiti, N. H. Kim and J. H. Lee, Strongly stabilized integrated bimetallic oxide of Fe2O3-MoO3 nano-crystal entrapped N-doped graphene as a superior oxygen reduction reaction electrocatalyst, Chem. Eng. J., 2021, 410, 128358 CrossRef CAS .
- Q. Dong, S. Ji, H. Wang, V. Linkov and R. Wang, Oxygen spillover effect at Cu/Fe2O3 heterointerfaces to enhance oxygen electrocatalytic reactions for rechargeable Zn-air batteries, ACS Appl. Mater. Interfaces, 2022, 14(45), 51222–51233 CrossRef CAS PubMed .
- Z. Yao, Y. Li, D. Chen, Y. Zhang, X. Bao, J. Wang and Q. Zhong, γ-Fe2O3 clusters embedded in 1D porous N-doped carbon matrix as pH-universal electrocatalyst for enhanced oxygen reduction reaction, Chem. Eng. J., 2021, 415, 129033 CrossRef CAS .
- Q. Yu, S. Lian, J. Li, R. Yu, S. Xi, J. Wu, D. Zhao, L. Mai and L. Zhou, FeNxand γ-Fe2O3co-functionalized hollow graphitic carbon nanofibers for efficient oxygen reduction in an alkaline medium, J. Mater. Chem., 2020, A8(12), 6076–6082 RSC .
- B. Zhang, T. Li, L. Huang, Y. Ren, D. Sun, H. Pang, J. Yang, L. Xu and Y. Tang, In situ immobilization of Fe/Fe3C/Fe2O3 hollow hetero-nanoparticles onto nitrogen-doped carbon nanotubes towards high-efficiency electrocatalytic oxygen reduction, Nanoscale, 2021, 13(10), 5400–5409 RSC .
- Q. Dong, H. Wang, J. Ren, X. Wang and R. Wang, Activating Cu/Fe2O3 nanoislands rooted on N-rich porous carbon nanosheets via the Mott-Schottky effect for rechargeable Zn-air battery, Chem. Eng. J., 2022, 442, 136128 CrossRef CAS .
- Y. Niu, Y. Yuan, Q. Zhang, F. Chang, L. Yang, Z. Chen and Z. Bai, Morphology-controlled synthesis of metal-organic frameworks derived lattice plane-altered iron oxide for efficient trifunctional electrocatalysts, Nano Energy, 2021, 82, 105699 CrossRef CAS .
- F. Zhang, Y. Zhu, Y. Zhong, J. Zou, Y. Chen, L. Zu, Z. Wang, J. J. Hinsch, Y. Wang, L. Zhang, Z. Shao and H. Wang, Tuning the charge distribution and crystal field of iron single atoms via iron oxide integration for enhanced oxygen reduction reaction in zinc-air batteries, J. Energy Chem., 2023, 85, 154–163 CrossRef CAS .
- X. Guo, S. Liu, X. Wan, J. Zhang, Y. Liu, X. Zheng, Q. Kong and Z. Jin, Controllable solid-phase fabrication of an Fe2O3/Fe5C2/Fe-N-C electrocatalyst toward optimizing the oxygen reduction reaction in zinc-air batteries, Nano Lett., 2022, 22(12), 4879–4887 CrossRef CAS PubMed .
- W. Yan, Y. Wu, Y. Chen, Q. Liu, K. Wang, N. Cao, F. Dai, X. Li and J. Jiang, Facile preparation of N-doped corncob-derived carbon nanofiber efficiently encapsulating Fe2O3 nanocrystals towards high ORR electrocatalytic activity, J. Energy Chem., 2020, 44, 121–130 CrossRef .
- X. Meng, J. Xie, Y. Sun, J. Liu, B. Liu, R. Wang, F. Ma, M. Liu and J. Zou, Fe2O3/spinel NiFe2O4 heterojunctions in situ wrapped by one-dimensional porous carbon nanofibers for boosting oxygen evolution/reduction reactions, Int. J. Hydrogen Energy, 2022, 47(50), 21329–21343 CrossRef CAS .
- Q. Xie, M. Pan, Z. Wang, W. Si, R. Zhang, Y. Shu, G. Sun, Q. Jing, Y. Shen and H. Uyama, Enhancing the oxygen reduction activity by constructing nanocluster-scaled Fe2O3/Cu interfaces, Nanoscale, 2023, 15(9), 4388–4396 RSC .
- R. Dun, M. Hao, Y. Su and W. Li, Fe-N-doped hierarchical mesoporous carbon nanomaterials as efficient catalysts for oxygen reduction in both acidic and alkaline media, J. Mater. Chem. A, 2019, 7(20), 12518–12525 RSC .
- B. J. Park, J. Kim and H. S. Park, Bifunctional electrocatalysts based on hierarchical graphene/iron hybrid architectures branched by N-doped CNT, J. Alloys Compd., 2020, 846, 156244 CrossRef CAS .
- Z. Xiao, C. Wu, W. Wang, L. Pan, J. Zou, L. Wang, X. Zhang and G. Li, Tailoring the hetero-structure of iron oxides in the framework of nitrogen doped carbon for the oxygen reduction reaction and zinc-air batteries, J. Mater. Chem. A, 2020, 8(48), 25791–25804 RSC .
- Q. Qian, H. Hu, S. Huang, Y. Li, L. Lin, F. Duan, H. Zhu, M. Du and S. Lu, Versatile hyper-cross-linked polymer derived porous carbon nanotubes with tailored selectivity for oxygen reduction reaction, Carbon, 2023, 202, 81–89 CrossRef CAS .
- C. Maouche, Y. Wang, C. Cheng, W. Wang, Y. Li, W. A. Qureshi, P. Huang, A. Amjad, Y. Zhou and J. Yang, Sulfur doped Fe-N-C catalysts derived from Dual-Ligand zeolitic imidazolate framework for the oxygen reduction reaction, J. Colloid Interface Sci., 2022, 623, 146–154 CrossRef CAS PubMed .
- R. Nandan, G. Raj and K. K. Nanda, FeCoNiMnCr high-entropy alloy nanoparticle-grafted NCNTs with promising performance in the ohmic polarization region of fuel cells, ACS Appl. Mater. Interfaces, 2022, 14(14), 16108–16116 CrossRef CAS PubMed .
- M. Shen, X. Lin, W. Xi, X. Yin, B. Gao, L. He, Y. Zheng and B. Lin, Mesoporous waffle-like N-doped carbon with embedded Co nanoparticles for efficiently electrocatalytic oxygen reduction and evolution, J. Colloid Interface Sci., 2023, 633, 374–382 CrossRef CAS PubMed .
- J. Zhao, Q. Li, Q. Zhang and R. Liu, Carbon tube-graphene heterostructure with different N-doping configurations induces an electrochemically active-active interface for efficient oxygen electrocatalysis, Chem. Eng. J., 2022, 431, 133730 CrossRef CAS .
- Z. Zhang, Y. Wang, J. Guan, T. Zhang, P. Li, Y. Hao, L. Duan, Z. Niu and J. Liu, Direct conversion of solid g-C3N4 into metal-ended N-doped carbon nanotubes for rechargeable Zn-air batteries, Inorg. Chem. Front., 2022, 9(14), 3428–3435 RSC .
- X. Zhang, Z. Zhao, X. Kong, H. Xu and W. Jin, Rational design of MOFs-derived Co-Ru species embedded N-doped carbon/carbon matrix for highly-efficient and multifunctional electrocatalysis, Appl. Surf. Sci., 2022, 606, 154818 CrossRef CAS .
- G. Ye, T. Zeng, S. Wang, S. Liu and Z. He, Etching-directed nitrogen doping strategy to construct efficient defective sites in carbon for electrocatalytic oxygen reduction, J. Power Sources, 2022, 549, 232122 CrossRef CAS .
- W. Zhang, M. Guillén-Soler, S. Moreno-Da Silva, A. López-Moreno, L. R. González, M. D. C. Giménez-López and E. M. Pérez, Mechanical interlocking of SWNTs with N-rich macrocycles for efficient ORR electrocatalysis, Chem. Sci., 2022, 13(33), 9706–9712 RSC .
- B. Liu, B. Yuan, C. Wang, S. You, J. Liu, X. Meng, X. Xu, Z. Cai, J. Xie and J. Zou, Highly-dispersed NiFe alloys in-situ anchored on outer surface of Co, N co–doped carbon nanotubes with enhanced stability for oxygen electrocatalysis, J. Colloid Interface Sci., 2023, 635, 208–220 CrossRef CAS PubMed .
- H. Wang, S. Su, T. Yu, C. Meng, H. Zhou, W. Zhao, S. Yan, T. Bian and A. Yuan, FeNi/NiFe2O4 hybrids confined in N-doped carbon sponge derived from Hofmann-type MOFs for oxygen electrocatalysis, Appl. Surf. Sci., 2022, 596, 153522 CrossRef CAS .
- B. Wang, K. Srinivas, Y. Liu, D. Liu, X. Zhang, W. Zhang and Y. Chen, N-doped CNTs capped with carbon layer armored CoFe alloy as highly stable bifunctional catalyst for oxygen electrocatalysis, Nano Res., 2022, 15(5), 3971–3979 CrossRef CAS .
- X. Cheng, Y. Wang, Y. Lu, L. Zheng, S. Sun, H. Li, G. Chen and J. Zhang, Single-atom alloy with Pt-Co dual sites as an efficient electrocatalyst for oxygen reduction reaction, Appl. Catal., B, 2022, 306, 121112 CrossRef CAS .
- W. Liu, C. Zhang, J. Zhang, X. Huang, M. Song, J. Li, F. He, H. Yang, J. Zhang and D. Wang, Tuning the atomic configuration of Co-N-C electrocatalyst enables highly-selective H2O2 production in acidic media, Appl. Catal., B, 2022, 310, 121312 CrossRef CAS .
- J. Zheng, T. Kang, B. Liu, P. Wang, H. Li and M. Yang, N-doped carbon nanotubes encapsulated with FeNi nanoparticles derived from defect-rich, molecule-doped 3D g-C3N4 as an efficient bifunctional electrocatalyst for rechargeable zinc-air batteries, J. Mater. Chem. A, 2022, 10(18), 9911–9921 RSC .
- K. Kim, K. Min, Y. Go, Y. Lee, S. E. Shim, D. Lim and S.-H. Baeck, FeCo alloy nanoparticles embedded in N-doped carbon supported on highly defective ketjenblack as effective bifunctional electrocatalysts for rechargeable Zn-air batteries, Appl. Catal., B, 2022, 315, 121501 CrossRef CAS .
- X. Yao, X. Wang, L. Sun, L. Li, E. Kan, B. Ouyang and W. Zhang, Popcorn-like Co3O4 nanoparticles confined in a three-dimensional hierarchical N-doped carbon nanotube network as a highly-efficient trifunctional electrocatalyst for zinc–air batteries and water splitting devices, Inorg. Chem. Front., 2022, 9(11), 2517–2529 RSC .
- Y. Zheng, X. Xu, J. Chen and Q. Wang, Surface O2-regulation on POM electrocatalyst to achieve accurate 2e/4e− ORR control for H2O2 production and Zn-air battery assemble, Appl. Catal., B, 2021, 285, 119788 CrossRef CAS .
- H. L. Zhao, S. Wu, C. Y. Liu, X. T. Yan, X. Xu, S. S. Fu, Y. B. Wang, Q. Su, X. Wang and Q. L. Yang, Encapsulation Fe-Nx combined with Co@C to construct efficient oxygen reduction catalysts with bimetallic sites and the application of Zn-air batteries, Mater. Today Chem., 2022, 26, 101174 CrossRef CAS .
- L. Song, H. Fan, X. Fan, H. Gong, T. Wang and J. He, A simultaneous phosphorization and carbonization strategy to synthesize a defective Co2P/doped-CNTs composite for bifunctional oxygen electrocatalysis, Chem. Eng. J., 2022, 435, 134612 CrossRef CAS .
- Y. Pan, X. Wang, H. Lin, Q. Xia, M. Jing, W. Yuan and C. M. Li, Three-dimensional Ni foam supported NiCoO2@Co3O4 nanowire-on-nanosheet arrays with rich oxygen vacancies as superior bifunctional catalytic electrodes for overall water splitting, Nanoscale, 2023, 15(34), 14068–14080 RSC .
- B. Liu, S. Wang, R. Feng, Y. Ni, F. Song and Q. Liu, Anchoring bimetal single atoms and alloys on N-doping-carbon nanofiber networks for an efficient oxygen reduction reaction and zinc-air batteries, ACS Appl. Mater. Interfaces, 2022, 14(34), 38739–38749 CrossRef CAS PubMed .
- X. Cui, Y. Tao, X. Xu and G. Yang, P-doping optimized d-band center position in MoO2 with enhanced oxygen reduction reaction and oxygen evolution reaction activities for rechargeable Zn-air battery, J. Power Sources, 2023, 557, 232519 CrossRef CAS .
- W. Xu, X. Zhang, X. Xu, J. Chen and Q. Wang, Guest molecule insertion-optimized d-band center position in MoS2 with improved sulfite activation ability inspired by sulfite oxidase, ACS Appl. Mater. Interfaces, 2023, 15(10), 13042–13051 CrossRef CAS PubMed .
- W. Xu, Y. Zhang, X. Zhang, X. Xu and Q. Wang, One stone, two birds: A Cu-S cluster as a laccase-mimicking nanozyme and sulfite activator for phenol remediation in marine environments, J. Hazard. Mater., 2023, 457, 131776 CrossRef CAS PubMed .
Footnote |
† Electronic supplementary information (ESI) available. Materials and characterization; RRDE test in ORR; RDE test in ORR; OER test; Zn–air battery assemble; theoretical calculation details; comparison ORR, OER activities of Ni/Fe-Fe2O3@NCNTs with other materials; the reaction routes in ORR and OER; XPS survey spectra of Ni/Fe-Fe2O3@NCNTs and Fe-Fe2O3@NCNTs; XPS high resolution C 1s spectrum of Ni/Fe-Fe2O3@NCNTs; SEM image of Ni/Fe-Fe2O3@NCNTs; SEM image of MIL-88A; CV of Ni/Fe-Fe2O3@NCNTs at different scanning rates; CV curves of Fe-Fe2O3@NCNTs under different scanning rates; LSV curves of Fe-Fe2O3@NCNTs under different rotating rates; K–L equation simulation for Fe-Fe2O3@NCNTs; n values of Fe-Fe2O3@NCNTs under different potentials; RRDE voltammograms of Fe-Fe2O3@NCNTs; n of Fe-Fe2O3@NCNTs; H2O2 selectivity of Fe-Fe2O3@NCNTs; CV curves of Ni/Fe-Fe2O3@NCNTs and Pt/C in PBS (0.1 M PBS); K–L equation simulation for Fe-Fe2O3@NCNTs in PBS (0.1 M PBS); H2O2 selectivity of Fe-Fe2O3@NCNTs in PBS (0.1 M PBS); chronoamperometric curve of Ni/Fe-Fe2O3@NCNTs in PBS (0.1 M PBS). See DOI: https://doi.org/10.1039/d4dt01733a |
|
This journal is © The Royal Society of Chemistry 2024 |