Yefei
Ma‡
ab,
Qiushi
Wang‡
a,
Xia
Xiao
a,
Zhong-Jie
Jiang
*c,
Weiheng
Chen
d,
Xiaoning
Tian
*b and
Zhongqing
Jiang
*a
aDepartment of Physics, Zhejiang Sci-Tech University, Hangzhou 310018, P. R. China. E-mail: zhongqingjiang@zstu.edu.cn
bDepartment of Materials and Chemical Engineering, Ningbo University of Technology, Ningbo 315211, P.R. China
cGuangzhou Key Laboratory for Surface Chemistry of Energy Materials, New Energy Research Institute & Guangdong Engineering and Technology Research Center for Surface Chemistry of Energy Materials, College of Environment and Energy, South China University of Technology, Guangzhou 510006, P. R. China. E-mail: eszjiang@scut.edu.cn
dVehicle Energy and Safety Laboratory, Department of Mechanical Engineering, Ningbo University of Technology, Ningbo 315336, P. R. China
First published on 4th December 2024
Although the use of transition metals as bifunctional catalysts for zinc–air batteries (ZABs) has obvious economic advantages, their performance in ZABs still fails to meet expectations due to the uncontrollable loading caused by the rapid nucleation rate of transition metals. In this study, controllable loading of an Fe/Co alloy on heteroatom-doped hollow graphene spheres (FeCo@NGHS) was realized via the regulation of small molecules. Sodium citrate, which served as a metal complexing agent and reaction buffer, effectively suppressed the excessive loading of Fe/Co alloy particles and facilitated the formation of Fe(Co)Nx active sites. Melamine, which served as a precursor for doping N atoms, provided anchor points for the loading of Fe/Co alloy particles and participated in the generation of Fe(Co)Nx. The fabricated catalyst had active sites with different chemical structures, such as pyridine-N, graphite-N, Fe(Co)Nx and Fe/Co alloy particles, all of which benefit the improvement of the oxygen reduction reaction/oxygen evolution reaction (ORR/OER) performance. Results showed that the fabricated FeCo@NGHS, which possesses the appropriate amount of Fe/Co alloy particles combined with the highest amount of formed Fe(Co)Nx active sites, exhibited the best ORR/OER bifunctional catalytic performance in alkaline electrolytes and excellent electrocatalytic stability. The ORR onset potential and half-wave potential were 0.961 V and 0.846 V (vs. RHE), respectively. The OER could achieve a low overpotential level of 391 mV at a current density of 10 mA cm−2. Furthermore, the rechargeable liquid ZAB and flexible all-solid-state (ASS) ZAB assembled by FeCo@NGHS exhibited higher discharge power density and longer charge–discharge cycle performance. FeCo@NGHS-based air cathodes exhibited outstanding performance in flexible ASS–ZABs, showing high open circuit voltage (1.45 V) and peak power density (74.06 mW cm−2). Thus, in clean energy storage and conversion technologies, a new synthetic strategy for constructing excellent bifunctional oxygen electrocatalysts is proposed in this work.
Transition metal-based materials have garnered attention in recent years for their impressive electrocatalytic performance and durability, making them promising candidates for various electrochemical applications.16–20 The raw materials of transition metals (such as Fe and Co) are all abundant and inexpensive.21,22 On the one hand, Fe-based catalysts can show excellent activity in ORR, while iron nitride-based carbon materials can even be comparable to the Pt-based catalysts.23 On the other hand, Co-based catalysts are highly promising in the OER active catalysts, including Co metal, cobalt oxide, and cobalt phosphide,24,25etc. Therefore, the combination of Fe and Co is expected to be an efficient and low-cost bifunctional catalyst. However, nanoparticles (NPs) of the above metals suffer from dissociation, migration, and aggregation problems, resulting in the decay of catalyst activity and stability. Researchers usually construct structural defects and introduce heteroatoms to anchor metal NPs on the surface of carbon-based materials, and also use pore-framing NPs as a way to achieve effective control of NPs.26–28 In addition, the low conductivity of the transition metals-based materials also affects their catalytic performance. Considering the need for material conductivity in electrocatalysis, catalyst substrates are often chosen from carbon nanotubes, graphene, etc.29–31 Except the transition metal, metal–N–C (metal = Fe, Co, Ni) has received much attention due to its high activity, which is centered on a metal–nitrogen mimicking a bioporphyrin center coordination structure. Researchers have reported on many excellent performing M–N–C electrocatalysts.32–35 The group of Kurungot prepared an Fe/Co-rich nitrogen-doped active site (Fe-/Co-NpGr) nanoporous graphene for ORR catalysis.36 Through a carefully controlled oxidative etching procedure, graphene was transformed into porous graphene (pGr). The resulting edge sites were used to dope nitrogen, and thus construct Fe/Co coordination centers (Fe/Co-NpGr) on the basis of nitrogen doping. X-ray photoelectron spectroscopy (XPS) and time-of-flight secondary mass spectrometry (ToF-SIMS) analyses revealed structural insights, confirming the existence of the M–N (M = Fe, Co) doped carbon active site ligands. The results show that the bimetallic ligand locked at the hole opening can significantly reduce the overpotential of the ORR during the reaction with the assistance of nitrogen.
Herein, we report a facile and environmentally friendly one-step method for the controllable loading of an Fe/Co alloy on heteroatom-doped hollow graphene spheres (FeCo@NGHS), which can be realized by small molecule regulation and effectively avoid the agglomeration of transition metal particles (as shown in Fig. 1). In this work, graphene hollow spheres were used as the carbon matrix due to the good conductivity, high specific surface area and excellent flexibility. Inspired by the synthesis process of Prussian blue analogues (PBA), sodium citrate was used as the metal complexing agent and reaction buffer for the controllable loading of Fe/Co alloy. Because the Co2+ ions can coordinate with sodium citrate, which is a competitor of [Fe(CN)6]3−, to moderate the proliferation rates during the nucleation and subsequent growth phases. According to the documented research, the synthesis of Co–Fe PBA proceeds at a swift pace. Without the addition of sodium citrate, nucleation takes place immediately after the direct mixing of Co2+ and [Fe(CN)6]3−, which generates a large number of nuclei and particles. The coordination of Co2+ ions with sodium citrate can slow the reaction kinetics, thus leading to the controllable loading of the Fe/Co alloy. Moreover, cobalt acetate and potassium ferricyanide work as the cobalt and iron sources, respectively. Nitrogen atoms can be introduced into the NGHS mainly with the help of melamine. The doping of nitrogen atoms with different chemical structure and the riveting of the Fe/Co alloy are closely related to the introduced sodium citrate and melamine. The designed catalyst FeCo@NGHS shows excellent ORR and OER performance, where the loaded Fe/Co alloy combined with heteroatom active sites, such as pyridine-N, graphite-N, and Fe(Co)Nx, all played an active role in the enhancement of the bifunctionality. The rechargeable liquid ZAB and flexible all-solid-state (ASS) ZAB assembled from FeCo@NGHS exhibit high discharge power density and more durable charge–discharge cycling performance. The FeCo@NGHS-based air cathode has outstanding performance in the flexible ASS–ZAB, presenting high open-circuit voltage (OCV) (1.45 V) and strong peak power density (74.06 mW cm−2).
Due to the complexation of sodium citrate,37 cobalt ions react with potassium ferricyanide slowly. After annealing, the sample FeCo@NGHS was fabricated. And the SEM image of FeCo@NGHS is shown in Fig. 2a. Furthermore, the uniform distribution of elements N, Fe and Co can be clearly seen from the EDS elemental mapping image of FeCo@NGHS (Fig. 2j). In order to explore the specific effect of sodium citrate, a controlled experiment was carried out. The preparation process was similar to that of sample FeCo@NGHS, but without the addition of sodium citrate, and the resulting sample was named as Fe&Co@NGHS. Fig. 2d shows an obvious particle aggregation in sample Fe&Co@NGHS compared with FeCo@NGHS. The total load of metal in sample Fe&Co@NGHS (7.30 wt%) is also higher than that in sample FeCo@NGHS (4.68 wt%) (Table S1†). It can be assumed that without the introduction of sodium citrate, more metal particles are anchored to the base material. The lack of complexation between sodium citrate and cobalt ions leads to the rapid complexation of cobalt ions with ferricyanide ions, thus forming more alloy particles.
![]() | ||
Fig. 2 SEM, TEM and HRTEM images of (a–c) FeCo@NGHS, (d–f) Fe&Co@NGHS, and (g–i) FeCo@GHS. (j) EDS element mapping image of FeCo@NGHS. The inset of (b) is a SAED pattern. |
Moreover, a comparative sample was designed and prepared to further understand the effect of melamine on the synthesized catalyst material. Sample FeCo@GHS was obtained by a preparation process similar to that of sample FeCo@NGHS without the introduction of melamine. Fig. 2g shows that without the introduction of melamine, the spherical structure in sample FeCo@GHS is less obvious than that in FeCo@NGHS and Fe&Co@NGHS, which is due to the lack of intercalation of melamine into GO sheets during the preparation process. The total load of metal (3.63 wt%) in sample FeCo@GHS is also lower than those of samples FeCo@NGHS and Fe&Co@NGHS (Table S1†). The inclusion of melamine is intimately linked to the incorporation of the alloy, as inferred from the analysis. Therefore, we can infer that melamine has a riveting effect on metal particles, and related studies have also suggested the anchoring effect of the nitrogen source urea on metal particles.38 Due to the presence of melamine, samples Fe@NGHS and Co@NGHS maintain a good spherical structure, as shown in Fig. S1a and c.† However, the load of a single metal for sample Fe@NGHS and Co@NGHS is about 0.82 wt% and 0.87 wt%, respectively, which is significantly reduced compared with sample FeCo@NGHS. Therefore, it can be supposed that under the same preparation conditions, compared with the single metal, the selected base material NGHS is more favorable for alloy anchoring.
![]() | ||
Fig. 3 (a) XRD patterns, (b) TGA curves, (c) Raman spectra, and (d) N2 adsorption/desorption isotherm curves, and the inset is the pore size distribution. |
A distinctive peak at 2θ = 22°–30° is consistently observed across all samples, which is attributed to the (002) lattice plane of graphite carbon (JCPDS # 41-1487) and stems from the carbonaceous substrates NGHS or GHS.40 High-resolution TEM (HRTEM) images in Fig. 2c, f, and i distinctly reveal lattice fringes consistent with the (002) plane of graphite. Different from other samples, Fe@NGHS has three diffraction peaks corresponding to graphite. The diffraction peak positioned at approximately 44° is attributed to the (100) plane of graphite carbon (JCPDS # 41-1487). Furthermore, an additional diffraction peak at approximately 15° is associated with the (001) plane of graphene carbon, indicating an interlayer spacing of 0.59 nm, a value derived from the Bragg's equation. The extended layer spacing compared with graphite results from the intercalation of the substance in the graphene oxide carbon sheets, such as PS spheres, melamine, sodium citrate and transition metal salts. Diffraction peaks at different positions indicate different carbon layer spacing, suggesting that the arrangement of carbon sheets in sample Fe@NGHS is the most disordered.
Fig. 3b shows the thermogravimetric analysis (TGA) of samples FeCo@NGHS, Fe&Co@NGHS and FeCo@GHS. Before the temperature reaches 350 °C, the mass percentage of the fabricated samples shows no significant decline, indicating the good thermal stability. When the temperature increases to 350–420 °C, significant material decomposition is found, which is due to the loss of the carbon substrate material at high temperature. After calcination in air at 800 °C, sample Fe&Co@NGHS shows the highest mass residue, suggesting the highest metal loading. In order to accurately measure the amount of loaded metal in the resulting samples, inductively coupled plasma mass spectrometry (ICP-MS) detection is adopted, and the result is listed in Table S1.† The total amount of loaded Fe/Co alloy in sample Fe&Co@NGHS is the highest, followed by sample FeCo@NGHS, and sample FeCo@GHS exhibits the lowest. The ICP-MS test results are consistent with the TGA results. It is interesting that the introduction of melamine is directly related to the amount of Fe/Co alloy. Another more important problem is that under the same conditions, the loading of the Fe/Co alloy is easier to achieve than that of a single metal, such as Fe or Co.
The Raman spectra of the resultant samples are shown in Fig. 3c, and all samples present the D and G bands at 1355 and 1584 cm−1, respectively.41 The D and G bands are symbols of the degree of disordered and graphitization structure of the material, where the D peak intensity is related to the defect site (sp3-hybridized carbon), while the G peak intensity is related to the amount of sp2-hybridized carbon in the carbon matrix.38 It is common to consider ID/IG as an evaluation parameter for the degree of graphitization of carbon materials.42 It is not difficult to find that compared with NGHS, the introduction of Co or the Fe/Co alloy significantly reduces the value of ID/IG for the resulting samples (such as FeCo@NGHS, Co@NGHS, Fe&Co@NGHS and FeCo@GHS), except for sample Fe@NGHS. The ID/IG value in sample Fe@NGHS increased significantly, consistent with the XRD result. It can be concluded that the introduction of Fe has a significant impact on the substrate material NGHS. A smaller value of ID/IG indicates that the carbon sheets are arranged in a more ordered manner, which signifies a heightened level of graphitization, resulting in improved electrical conductivity. At the same time, a higher value of ID/IG suggests the presence of more defect sites in the prepared sample.
Fig. 3d depicts the N2 sorption/desorption isotherms and the corresponding pore size distribution profiles for the prepared samples. Compared with NGHS, the specific surface area of the obtained catalyst material is not significantly reduced after the metal loading. On the contrary, some of them even improved significantly, which should be caused by the insertion of sodium citrate and transition metal salts. Especially, FeCo@NGHS presents the highest specific surface area (617.4 m2 g−1), indicating the exposure of more active sites. As shown in Fig. 3d, all samples show a distinct type IV isotherm in the medium pressure region (P/P0 = 0.5–1.0) with a sharp increase in nitrogen uptake at relatively low pressures,43 which reveal the presence of both microporous and mesoporous structures in the material. These findings are in line with the pore size distribution data presented in the inset. The high specific surface area and rich micro/mesoporous structure of the fabricated catalyst can provide higher mass transport speed for the electrocatalytic reaction, while the microporosity accelerates the electron transfer,44 which effectively enhances the electrocatalytic activity with the synergy of the high specific surface and micro/mesopores.
The XPS wide-scan full spectrum in Fig. 4a shows the elemental composition on the surface of the resulting catalyst materials. The high-resolution C 1s XPS spectra are shown in Fig. 4b. Each spectral line can be resolved into four distinct peaks, corresponding to C–C at 284.8 eV, CN and C–O at 285.85 eV, C
O and C–N at 288.75 eV, and O–C
O at 292.4 eV, respectively.45 Furthermore, the sp2 hybridized C–C is the main component, which can suggest the graphite structure of substrate NGHS and indicate the good conductivity. Fig. 4c directly displays the successful doping of N atoms into the carbon skeleton.46 Melamine serves as the main nitrogen precursor for the doping of N atoms. The doped N atoms, which are strongly electronegative, can cause charge rearrangement and result in more active sites. In the high-resolution N 1s XPS spectrum of FeCo@NGHS (Fig. 4c), it can be deconvoluted into five peaks at 398.45 eV, 399.38 eV, 400.14 eV, 400.98 eV, and 402.98 eV, which are attributed to pyridine-N (398.45 eV), Fe–/Co–N (399.38 eV), pyrrole-N (400.14 eV), graphite-N (400.98 eV), and oxidized-N (402.98 eV), respectively.47 Furthermore, all control samples, except for NGHS and FeCo@GHS, contained a metal–N active site.48 Additionally, as can be observed from Table S2,† the sample FeCo@NGHS exhibits the highest content of the Fe–/Co–N component.
According to Fig. 4a, all samples exhibit high nitrogen doping, except the sample FeCo@GHS, which is synthesized without the introduction of melamine. From Fig. 4c, it can be found that pyridine-N and graphite-N are the main structures of nitrogen doping in the obtained samples, and they can promote the positive shift of the ORR onset potential and increase the diffusion-limited current density, respectively.49,50 Metal nitrides (Fe(Co)Nx) also show positive correlation properties with electrocatalytic performance. As shown in Fig. 4c and d, the sample FeCo@NGHS contains higher metal nitrides (Fe(Co)Nx) and adsorbed oxygen content than other control samples, which can promote the effect of the bifunctional catalysis (Table S3†).
The introduction of melamine is directly related to the doping of N, and the doping of N atoms can not only introduce the active site of N atoms, but also provide anchor points for metal loading. Comparing the metal loading in FeCo@GHS and FeCo@NGHS, it can be seen that the absence of melamine leads to a significant decrease in alloy loading (Table S1†). In addition, compared to FeCo@NGHS, the lack of sodium citrate in Fe&Co@NGHS leads to rapid nucleation and excessive loading of the alloy. Although the loading of the alloy in Fe&Co@NGHS is high, the amount of Fe(Co)Nx obtained is obviously lower than FeCo@NGHS. Therefore, it can be inferred that the complexation of sodium citrate with cobalt ions slows down the formation of alloy particles, thus promoting the generation of more Fe(Co)Nx active sites. The FeCo@GHS sample has fewer N and Fe(Co)Nx active sites compared to FeCo@NGHS due to the absence of melamine during its synthesis process, which is the main reason for its poor electrocatalytic performance.
As discussed above, the introduction of sodium citrate can achieve the controllable loading of alloy particles, and the riveting of the metal by doped N resulted in a moderate amount of metal loading in sample FeCo@NGHS, accompanied by the formation of a large number of Fe(Co)Nx active sites. Combining the above analytical results, it is easy to see that the introduction of sodium citrate and melamine is the key to form more Fe(Co)Nx active sites. It is interesting that metal–Nx active sites, such as Fe–Nx and Co–Nx types, have been confirmed by the previously reported research work for their strong coupling effect between N-doped graphene shells and metal NPs, accelerating the mass transfer efficiency and thus enhancing the catalytic performance.51 Therefore it can be assumed that the formation of Fe(Co)Nx is related to both melamine and sodium citrate, and the absence of either will reduce the N content and the percentage of Fe(Co)Nx. From Fe@NGHS, it can be seen that the doping amount of N is relatively high, but the loading amount of Fe is low, indicating that the loading of a single metal is difficult under this synthesis condition.
As shown in the high-resolution XPS spectra of Fe 2p (Fig. 4e), five pairs of double peaks are present in sample FeCo@NGHS, with zero-valence Fe peaks observed at 707.73 eV (Fe 2p3/2) and 719.56 eV (Fe 2p1/2). This indicates the presence of metallic Fe, which can be attributed to metallic Fe in alloy Co0.7Fe0.3. A pair of main peaks is located at 711.8 eV (Fe 2p3/2) and 724.29 eV (Fe 2p1/2), while satellite peaks are observed at 716.4 eV (Fe 2p3/2) and 728.13 eV (Fe 2p1/2), originating from Fe2+/3+.39 The Fe 2p high-resolution XPS spectra of samples Fe@NGHS and Fe&Co@NGHS also contain five pairs of double peaks, but there is no double peak representing Fe(Co)Nx in sample FeCo@GHS. Similarly, the high-resolution XPS spectra of Co 2p (Fig. 4f) show the presence of zero-valence Co peaks at 777.1 eV (Co 2p3/2) and 792.9 eV (Co 2p1/2), attributable to metallic Co in the alloy Co0.7Fe0.3. Co–Nx peaks are located at 778.4 eV (Co 2p3/2) and 794.3 eV (Co 2p1/2), while the satellite peaks are located at 786.0 eV (Co 2p3/2) and 800.5 eV (Co 2p1/2), respectively, suggesting the presence of the Co–Nx active sites.39 Similarly, the Co 2p high-resolution XPS spectra of samples Co@NGHS and Fe&Co@NGHS include the same Co component, but the FeCo@GHS sample lacks the Co–Nx component. The high-resolution XPS spectra of the obtained sample FeCo@NGHS once again demonstrate the presence of Fe(Co)Nx active sites, as well as the metal and alloy composition in the loaded particles.
As shown in Fig. 5b, the EIS analysis of the samples FeCo@NGHS, Fe@NGHS, Co@NGHS, Fe&Co@NGHS, FeCo@GHS, NGHS, and 20 wt% Pt/C were performed. It can be observed that the corresponding semicircle diameters of FeCo@NGHS, Fe@NGHS, 20 wt% Pt/C, Co@NGHS, Fe&Co@NGHS, FeCo@GHS and NGHS gradually increase, showing an increase in the charge transfer resistance (Rct). FeCo@NGHS shows the smallest Rct of 39.96 Ω (Table S5†), exhibiting the best ORR activity.52 Similarly, Fig. 5c shows that the Tafel slope of FeCo@NGHS is 86.56 mV dec−1, which is lower than that of Fe@NGHS (157.8 mV dec−1), Fe&Co@NGHS (104.29 mV dec−1), FeCo@GHS (118.23 mV dec−1), NGHS (122.23 mV dec−1) and 20 wt% Pt/C (86.81 mV dec−1), followed only by 65.11 mV dec−1 from Co@NGHS. The lower Tafel slope indicates that the catalyst FeCo@NGHS has a reaction mechanism and kinetics similar to those of commercial Pt/C.53 Compared with other bimetallic loaded samples, it can be assumed that more active sites of metal–Nx–C are formed in it due to the highest total N doping and percentage of Fe(Co)Nx in sample FeCo@NGHS. Therefore, the lower Tafel slope of FeCo@NGHS may by caused by the synergistic interaction between metal NPs and nitrogenous material, which accelerates the dissociation of OOH* (OOH* → O* + OH*), further highlighting the structural advantage of M–N–C.54,55
From the CV curves in Fig. 5d, each catalyst exhibits a clear redox peak. The position of the oxygen reduction peak of the FeCo@NGHS (0.867 V) is higher than the potential of the rest of the control samples, such as Fe@NGHS (0.821 V), Co@NGHS (0.830 V), Fe&Co@NGHS (0.847 V), FeCo@GHS (0.810 V) and NGHS (0.741 V), which are consistent with the LSV comparison results. In the N2-saturated electrolyte, none of the CV curves of the samples have ORR peaks, indicating that none of them had Faraday reactions. This also further confirms that the sample FeCo@NGHS has outstanding ORR performance.
As mentioned in the introduction, the catalytic activity of ORR catalysts is usually measured by the number of electron transfers, so the samples were tested for RRDE (the results are shown in Fig. 5e). The obtained data were converted to obtain Fig. 5f, where the electron transfer numbers and hydrogen peroxide yields of FeCo@NGHS, Fe@NGHS and Fe&Co@NGHS are close to those of Pt/C, indicating that the electron transfer mechanism of the catalysts is consistent with the four-electron pathway of commercial catalysts. The K–L plots in Fig. S2† present the linear correlation of the designed electrode catalysts at different potentials, indicating that the electrode catalysts have the primary reaction kinetics.53
Through the comparison of the ORR performance, the ORR comprehensive performance order of the designed catalysts is as follows: FeCo@NGHS > Pt/C ≈ Co@NGHS > Fe&Co@NGHS > Fe@NGHS > FeCo@GHS > NGHS. FeCo@NGHS exhibits better performance than commercial Pt/C in terms of Rct, kinetic current density, etc., which indicates that FeCo@NGHS has the best kinetic performance. On the one hand, the sample FeCo@NGHS has a moderate alloy loading due to the riveting effect of the doped N atoms. With the introduction of sodium citrate, the alloying reaction of potassium ferricyanide with cobalt proceeded controllably under the complexation of cobalt ions by sodium citrate, resulting in the formation of more Fe(Co)Nx active sites in FeCo@NGHS, compared with Fe&Co@NGHS. On the other hand, the FeCo@GHS without melamine incorporation exhibits poor performance, similar to the substrate material NGHS. This is mainly due to the lack of N doping, resulting in the decrease of the alloy load and the absence of a sufficient number of effective Fe(Co)Nx active sites. Moreover, the ORR activities of the single-metal samples Fe@NGHS and Co@NGHS are both weaker than those of FeCo@NGHS catalysts, due to the synergistic effect of the Fe/Co alloy, which can promote the ORR catalysis.
Durability is also critical for ORR catalysts. Fig. 5g shows the chronoamperometry response tests of FeCo@NGHS and 20 wt% Pt/C at 0.85 V. It can be observed that the catalyst FeCo@NGHS has good durability. Its relative current can still be maintained at 80.76% after 2.7 h of testing, while the commercial Pt/C is maintained at 76.32%. The methanol oxidation resistance and CO toxicity resistance of the catalysts were also measured. As shown in Fig. 5i, the current density of FeCo@NGHS or 20 wt% Pt/C decreased after adding methanol or passing CO. However, FeCo@NGHS did not change drastically compared with the noble metal catalysts, confirming that it has better resistance to methanol oxidation and the anti-CO toxicity ability. It further illustrates its application potential in fuel cell systems. The good activity and durability of the catalyst originate from the protection by the graphite structure and carbon matrix. After continuous ORR operation, the graphite substrate of FeCo@NGHS remained unchanged. Fig. S3† shows the SEM and TEM results, where the complete graphene sphere morphology could be observed. Metal NPs are still dispersed and supported in NGHS without obvious aggregation. As shown in the SAED (inset of Fig. S3b†), the diffraction rings corresponding to the crystal planes (211), (200) and (110) of the Fe/Co alloy still exist. After the ORR stability test, the XPS evaluation of the sample FeCo@NGHS was conducted to analyze its surface element composition and electronic structure. It can be observed from Fig. S4† that the overall elemental composition of the FeCo@NGHS remains unchanged, and the electronic structures of Fe and Co elements on the surface also remain unchanged.
The CV tests for the catalysts were performed at different scan rates (20–100 mV s−1) in the faradaic region, and the Cdl data (shown in Fig. S5†) were obtained to evaluate the ECSA of the catalysts due to the fact that these two have a linear relationship.56 As shown in Fig. S5g,† the calculated Cdl values of FeCo@NGHS, Fe@NGHS, Co@NGHS, Fe&Co@NGHS, FeCo@GHS and NGHS are 11.43 mF cm−2, 11.19 mF cm−2, 8.88 mF cm−2, 10.2 mF cm−2, 5.97 mF cm−2 and 7.31 mF cm−2, respectively. The FeCo@NGHS exhibits the largest Cdl value, which indicates that FeCo@NGHS has a larger ECSA and can create more abundant active sites for the OER process.12
The performance of FeCo@NGHS is significantly better than that of Fe&Co@NGHS. This is due to the regulation of the reaction rate between potassium ferricyanide and cobalt ions by the complexing agent sodium citrate. This enables the sample FeCo@NGHS to have appropriate alloy loading and form a large number of highly active Fe(Co)Nx active sites, leading to better OER performance. Due to the lack of sufficient doped N atoms, the lowest content of alloy loading, N atom doping and formed Fe(Co)Nx active sites resulted in the worst OER performance of FeCo@GHS in the alloy-loaded samples. Furthermore, the OER activity and kinetics of FeCo@NGHS are superior to those of the single-metal samples Fe@NGHS and Co@NGHS, which originate from the synergistic effect of Fe/Co alloys.
In the chronopotentiometry test (as shown in Fig. 6d), the potential of FeCo@NGHS hardly changed after working continuously for 63 h in 0.1 M KOH electrolyte with a current density of 10 mA cm−2. Under the same conditions, the potential decay of RuO2 is obvious (≈77%). The corresponding morphology and elemental composition of FeCo@NGHS were characterized after 63 h of OER testing. As shown by the SEM and TEM in Fig. S3d–f,† the FeCo@NGHS sample still has the morphology of uniform graphene spheres, and there are well-dispersed alloy NPs. According to the SAED (inset of Fig. S3e†), the (110), (200) and (211) planes belong to the Fe/Co alloy still clearly visible.
The high bifunctional performance of the sample FeCo@NGHS originates from the controllable loading of the Fe/Co alloy: (1) the complexation of sodium citrate constrains the combination of cobalt ions and potassium ferricyanide, resulting in the formation of more Fe(Co)Nx active sites; (2) melamine plays the main role in N doping and the following riveting of the alloy particles. It also has a certain role in forming hollow graphene spheres with high specific surface area. For FeCo@NGHS, the doped N atoms with high electronegativity will affect the adjacent carbon atoms, presenting a higher density of positive charges. It facilitates the chemical adsorption of oxygen molecules and diminishes the strength of the O–O bond, thereby accelerating the rate of the ORR.57,58 In addition, the Fe(Co)Nx active sites, formed by the doped N atoms and the subsequent loading of the obtained alloy particles, play a key role in the improvement of the ORR and OER performance; (3) the material is based on graphene hollow spheres. The graphite substrate has high electronic conductivity and abundant structural defects. Graphene presents a three-dimensional spherical structure, which can accelerate the mass transfer of the reactants and reaction intermediates, and provide more active sites.
The bifunctionality of catalysts is typically assessed by quantifying the potential gap (ΔE) between the potential required for achieving 10 mA cm−2 during the OER process and the half-wave potential observed during the ORR reaction.59 The ΔE value of FeCo@NGHS is 0.775 V, which is lower than that of commercial Pt/C + RuO2 (0.87 V) (Fig. 5h). It proves that the catalyst FeCo@NGHS has excellent bifunctional activity, and the bifunctional activity of the sample is comparable to metal-based carbonaceous catalysts that have been previously reported in many studies (see Table S4†).
To investigate the cycling stability of the catalyst in the ZAB (as shown in Fig. 7e), the galvanostatic charge–discharge cycle test was performed at 10 mA cm−2. The catalyst Pt/C + RuO2 shows a voltage difference of 1.50 V after 210 cycles, while the energy efficiency decreased from 54.2% to 40.7%. The weakening of the stability may be due to the reduction of the interaction between Pt NPs and carbon materials, and the loss of RuO2 activity.39 Similarly, the electrocatalyst FeCo@NGHS-assembled ZAB only increases the voltage difference from the initial 1.01 V to 1.09 V after 210 cycles (∼33 h), and the energy efficiency only decreased from the initial 53.4% to 51.1%, which confirmed the excellent charge–discharge cycle stability. The superior performance could stem from the electrocatalyst's robust structural integrity and stable surface chemistry.
Based on the excellent performance of rechargeable liquid ZABs, a flexible ASS–ZAB was assembled based on FeCo@NGHS (inset of Fig. 8a). Fig. 8a shows the OCV (1.45 V) of the FeCo@NGHS-based ASS–ZABs, which is slightly higher than that of ASS–ZABs assembled with commercial Pt/C + RuO2 (1.42 V). The discharge polarization curves and the corresponding power density curves are presented in Fig. 8b, in which the ASS–ZAB assembled with FeCo@NGHS catalyst exhibits a peak power density of 74.06 mW cm−2. This is better than that of commercial Pt/C + RuO2 (35.35 mW cm−2) and many advanced electrocatalysts in ASS–ZABs (shown in Table S7†). Fig. 8c displays the galvanostatic charge–discharge cycling curve at 1 mA cm−2, with an initial discharge potential of 1.32 V and a charge potential of 1.96 V, respectively. The initial charge–discharge voltage difference of the battery is 0.64 V, which increases to 0.76 V by the 40th cycle, corresponding to a decrease in energy efficiency from the initial 67.3% to 62.5%. To further study the flexibility of the ASS–ZAB, the battery was bent at different angles and simultaneously tested its charge–discharge cycle stability at 1 mA cm−2, as shown in Fig. 8d. To be sure, the battery still maintains a certain discharge (1.28 V) and charge (1.95 V) plateau after being bent at any angle. In addition, the potentials gap and round-trip efficiency of the as-assembled battery show no significant change under various bending angle. Furthermore, the ASS–ZABs can exhibit a voltage retention of 98.3% when bended from initial 0° (1.448 V), 90° (1.431 V) to 180° (1.424 V), suggesting its good flexibility and stability.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qi02166e |
‡ These authors contributed equally. |
This journal is © the Partner Organisations 2025 |