Next Article in Journal
Mechanistic and Experimental Study of the CuxO@C Nanocomposite Derived from Cu3(BTC)2 for SO2 Removal
Previous Article in Journal
Hollow Nanospheres Organized by Ultra-Small CuFe2O4/C Subunits with Efficient Photo-Fenton-like Performance for Antibiotic Degradation and Cr(VI) Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of Sulfate-Intercalated CoFe LDH Nanosheets Derived from Two-Dimensional ZIF-9(III) for Promoted Oxygen Evolution Reaction

Guangxi Key Laboratory of Low Carbon Energy Materials, School of Chemistry and Pharmaceutical Sciences, Guangxi Normal University, Guilin 541004, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 26 May 2022 / Revised: 18 June 2022 / Accepted: 21 June 2022 / Published: 23 June 2022

Abstract

:
Layered double hydroxide (LDH) has emerged as a promising electrocatalyst; however, the synthetic method usually requires high temperature and high pressure, and sulfate-intercalated LDH is rarely reported. Herein, the sulfate-intercalated CoFe LDH nanosheets were successfully fabricated at ambient temperature via a facile strategy, using two-dimensional ZIF-9(III) as a template and FeSO4 as both etchant and iron source. When the as-prepared sulfate-intercalated CoFe LDH acts as an electrocatalyst, it presents superior electrocatalytic performance for the oxygen evolution reaction (OER), requiring low overpotential (η@10 mA cm−2 = 218 mV) with a small Tafel slope of 59.9 mV dec−1 in 1.0 M KOH, which compares favorably with commercial RuO2 and most reported transition-metal electrocatalysts. The high catalytic activity of CoFe LDH might be ascribed to the large interlayer space distance originating from special SO42− ions and the strong synergistic effects between Fe and Co. This work provides a novel and feasible approach to designing highly efficient electrocatalysts based on advanced LDH materials for OER.

Graphical Abstract

1. Introduction

Hydrogen energy has been extensively explored as an ideal green alternative to traditional fossil fuels because of its high energy density, perfect conversion efficiency, and zero carbon emissions [1]. Water electrolysis provides a potential path to produce high-purity hydrogen. Nonetheless, the efficiency of water electrolysis is greatly hampered by the sluggish kinetics of the oxygen evolution reaction (OER), which involves a complex multielectron and multiproton process. Currently, noble metal oxides (e.g., RuO2 and IrO2) are the benchmark of OER electrocatalysts, but their scarcity, prohibitively high cost, and unsatisfactory stability have severely limited their widespread commercial application [2,3]. Therefore, exploiting the high efficiency, large reserves, and low costs of non-precious-metal-based catalysts to accelerate the reaction rate of OER has become an imperative and urgent matter.
Transition-metal (e.g., Co, Fe and Ni) layered double hydroxides (LDH) have emerged as promising electrocatalysts for OER because of their flexible open layer structures, tunable chemical composition, and cost effectiveness [4,5,6]. In particular, the OER activity of bimetallic CoFe LDH is dramatically improved compared to the individual Co and Fe components due to the modulated electronic structures, enhanced charge transfer, and synergistic interactions between Co and Fe [7,8]. However, the bulk CoFe LDH still suffers from both intrinsically weak electronic conductivity and serious aggregation problems. To acquire high-performance CoFe LDH electrocatalysts, various strategies have been adopted, including heteroatom doping, the design of nanostructures, combining CoFe LDH with other active materials, and the exfoliation of CoFe LDH et al. [9,10,11,12]. For example, Wang’s group developed a plasma-enabled exfoliation strategy to obtain ultrathin CoFe-LDH nanosheets, which exhibited superior catalytic activity compared to the pristine samples because they were exposed to more active sites [12]. It is well known that LDH nanosheets have abundant active sites, rapid mass transport, and superior electron transfer ability, thereby greatly promoting catalytic properties. On the other hand, expanding the interlayer distance by changing interlayer anions is also proved to be an efficient strategy to create more active sites and improve electrocatalytic activity [13,14,15]. For example, Sun’s group found that more intercalated NO3 instead of CO32− could enlarge the interlayer space from 7.69 Å to 8.04 Å, thus reducing the charge-transfer resistance and enhancing the catalytic activity of NiFe LDH [16]. Furthermore, Lang’s group reported ultrathin sulfate-intercalated NiFe-LDH nanosheets that show excellent stability and great electrocatalytic performance in OER [17]. Therefore, developing a facile synthetic method to fabricate sulfate-intercalated CoFe LDH nanosheets with a large interlayer space is highly desirable.
Recently, due to their high surface area, dispersed metal centers, and abundant pore structures, metal–organic frameworks (MOFs) have been suggested as one of the best precursors for preparing the metal oxides/hydroxide or porous carbon materials through pyrolysis or etching [18,19]. Moreover, numerous studies have demonstrated that these MOFs derivatives are promising candidates for OER electrocatalysts [20,21,22]. In particular, most MOFs-derived LDHs exhibit excellent OER electrocatalytic performances [23,24,25]. It is worth noting that the Co-N bond in ZIF-67 structure can be easily broken in acidic conditions and that the released Co2+ ions can be used as metal precursors to build MOF-derived LDHs [26,27]. However, the microstructure of the involved precursor MOFs usually consists of three-dimensional (3D) nanomaterials and the transformation from 3D MOFs into 2D LDH is a high-energy-consuming and complicated synthetic process, leading to a high cost [28]. In our previous work, a novel 2D ZIF-9(III)/Co LDH was successfully synthesized via the controllable phase transition of 3D ZIF-9(I) and exhibited an efficient OER electrocatalytic performance [29]. Thus, developing a facile, surfactant-free, high-yield, and economically feasible synthetic route based on 2D ZIF-9 to fabricate CoFe LDHs with large interlayer space is a considerable challenge.
Herein, a facile method of etching–coprecipitation is developed to drive the conversion of 2D ZIF-9(III) into sulfate-intercalated CoFe LDH nanosheets. The as-prepared CoFe LDH not only retains the 2D morphology of the ZIF-9(III) precursor, but also exhibits a hierarchical structure composed of small crimped nanosheets. Benefiting from the larger interlayer spacing with sulfate intercalation, abundant active sites, and synergistic effect between Co and Fe, the CoFe LDH shows extraordinary mass transport and excellent activities for OER in alkaline media.

2. Results and Discussion

The synthetic process of the CoFe LDH nanosheets is schematically illustrated in Figure 1. Firstly, uniform 2D ZIF-9(III) nanosheets were fabricated using a simple precipitation method at room temperature. Then, sulfate-intercalated CoFe LDHs were successfully prepared through a gentle etching–coprecipitation method using FeSO4·7H2O as the etchant and 2D ZIF-9(III) as the precursor. During the chemical transformation, the protons generated from the hydrolysis of Fe2+ gradually attacked the Co-N bond of ZIF-9(III) to release the Co2+ ions into the solution. At the same time, Fe(OH)2 from the hydrolysis of Fe2+ was oxidized to form Fe(OH)3 in the air. Afterwards, Co2+ and Fe(OH)3 coprecipitated homogeneously to form CoFe LDH nanosheets on the outer layer of the ZIF-9 precursors, while the SO42− was reserved as intercalation anions. Finally, with the gradual dissolution of the precursor, hierarchical CoFe LDH composed of ultrathin nanosheets was obtained.
X-ray diffraction (XRD) analysis was used to investigate the composition and structural parameters of the as-synthetic samples. The characteristic XRD pattern of ZIF-9(III) was identified as the simulated XRD pattern of ZIF-9(III) (CCDC-988184, Figure 2a) [30], which indicates that the precursor was successfully prepared. As shown in Figure 2b, the XRD pattern of the CoFe LDH nanosheets could be assigned to a typical hydrotalcite-like structure [31], demonstrating the complete conversion of the ZIF-9(III) into CoFe LDH using the facile method. The four peaks were observed at the 2θ values of 10.0°, 20.3°, 33.4°, and 59.7°, which can be attributed to the (003), (006), (012), and (110) planes of LDH (JCPDS No. 52-0552), respectively. The LDH interlayer spacing was calculated using Bragg’s law: d = nλ/2sinθ. It should be pointed out that the selected 2θ values correspond to the (003) crystal plane from the structure of the LDH [15]. Thus, the interlayer spacing was calculated to be 0.89 nm from the XRD result (Figure 2b), which is larger than that of previously reported LDH [10,11,12]. As we know, increasing the interlayer spacing could accelerate the electrocatalytic activity for the water oxidation reaction [11,14].
To investigate a fingerprint of molecular vibrations of the bonding structure from the as-prepared samples, Fourier transform infrared spectroscopy (FTIR) was utilized. As shown in Figure 2c, the broad band located at 3430 cm−1 was attributed to the hydroxyl groups of the adsorbed water molecules. A broad band ranging from 500–650 cm−1 in the FTIR spectrum of ZIF-9(III) related to the stretching vibration of the Co-N bonds. The absorption at 1606 cm1 was attributed to the C=N stretching mode [32]. The peaks at about 745 cm1 and 1461 cm1 corresponded to the C-H out-of-plane bending vibration and the ring vibrations, respectively [33]. The above results demonstrate the successful preparation of ZIF-9(III). For CoFe LDH, the presence of an obvious and wide-ranging band at around 3430 cm−1 demonstrates the stretching vibration of hydrogen-bonded hydroxyl groups from both the brucite-like layers and the intercalated H2O molecules. The sharp peak at around 1629 cm−1 might originate from the bending vibration of water [34]. The weak bands below 800 cm−1 may belong to the characteristic bending vibration of Fe-O and Co-O. The bands at 1113, 1066, and 618 cm−1 [17,35,36] are ascribed to the vibrations of S-O, demonstrating that the main interlayer anion in the LDH is SO42−. The bivalent SO42− can be intercalated more easily into the interlayer of LDH due to the stronger electron affinity between the SO42− anions and the positively charged metal-OH host layers [15]. The Raman spectra of the ZIF-9(III) and CoFe-LDH are displayed in Figure 2d. All the bands of 2D ZIF-9(III) are found in the same position as in previously reported work [37]. In addition, the two obvious peaks at 446 and 526 cm−1 for CoFe LDH are ascribed to the M-O vibration [38]. The strongest intensity band at 981 cm−1 can be observed in the Raman spectrum of the CoFe LDH, which is assigned to the symmetric stretching of SO42− tetrahedra [39], in agreement with the results of FTIR.
Surface morphologies of nanostructured ZIF-9(III) and CoFe LDH were characterized using field emission scanning electron microscopy (FE-SEM) and transmission electron microscopy (TEM). As displayed in Figure 3a, the sheet-like morphology of ZIF-9(III) can be observed. The corrugations of some wrinkles in the sample indicate that the ZIF-9(III) is a multilayered structure (Figure S1). Moreover, the CoFe LDH maintains the size and morphology of the precursor in Figure 3b. However, the flat surface is covered by substantial and densely packed nanosheets with smaller lateral size and thickness. As shown in the high-magnification FE-SEM images (Figure 3c,d and Figure S2), the CoFe LDH is composed of many ultrathin nanosheets, indicating that the material displays a hierarchical structure. Furthermore, the structure of CoFe LDH is analyzed using TEM and HRTEM. As shown in Figure 3e,f, the CoFe LDH is composed of ultrathin nanosheets. The folded edges or wrinkles of nanosheets illustrate the ultrathin character. The atomic force microscopy (AFM) results (Figure S3) demonstrate that the average thickness of the ZIF-9(III) and CoFe LDH nanosheets is about 1.75 nm and 1 nm, respectively, and the lateral size is about 1.5 μm and 100–200 nm, respectively. Such thin nanosheets could provide rich redox-reaction sites to improve the electrochemical performance of CoFe LDH [2,12]. Figure 3g and Figure S4 demonstrate HRTEM images of the nanosheet, and the spacing of 0.890 nm, 0.265 nm, and 0.445 nm correspond to the (003), (012), and (006) plane of the LDH structure, respectively, in accordance with the above XRD data. The corresponding EDX mappings confirm the uniform distribution of Co, Fe, O, and S on the nanosheets (Figure 3h). In addition, the BET surface area and porosity of the CoFe LDH are characterized using N2 isothermal adsorption–desorption measurements (Figure S5). The surface area of CoFe LDH is 57.0 m2 g−1. From the corresponding Barrett–Joyner–Halenda (BJH) pore-size distributions, CoFe LDH exhibits the mesoporous property.
The surface compositions and the detailed valence states of samples were investigated using X-ray photoelectron spectroscopy (XPS) measurements. The binding energy corrections on high-resolution scans were calibrated by referencing the C 1s peak (284.8 eV). The survey spectra of CoFe LDH confirms the coexistence of Co 2p, Fe 2p, O1s, and S 2p (Figure 4a). As can be seen from the Co 2p spectrum (Figure 4b), the two peaks located at 781.2 and 797.3 eV are assigned to the Co 2p3/2 and Co 2p1/2 spin-orbit peaks, respectively. According to the peak-fitting analysis, the characteristic peaks of binding energy at 782.7 and 798.6 eV are ascribed to Co2+, while the other two peaks at 781.1 and 797.1 eV are attributed to Co3+ [13]. Additionally, the presence of the peaks at 786.8 and 803.4 eV are assigned to the typical satellite peaks of 2p3/2 and 2p1/2 spin orbits (denoted as “sat.”), respectively. Interestingly, compared with our previously reported Co 2p for Co LDH, peaks of both Co 2p3/2 and Co 2p1/2 for CoFe LDH have a slightly positive shift [29]. The result indicates that the electron cloud arrangement of the Co atom could be regulated by incorporating the Fe atom and the synergistic effect of the two atoms could be further justified. For the Fe 2p spectrum in Figure 4c, two prominent peaks located at 726.0 eV and 712.5 eV are attributed to Fe 2p1/2 and Fe 2p3/2, respectively, confirming that Fe2+ ions are oxidized to Fe3+ ions during the reaction process [10]. In the C 1s spectrum, peaks associated with adventitious carbon (284.8 eV), C-O (286 eV), and C=O (288.9 eV) moieties are observed. As shown in Figure 4e, it can be clearly seen that the high-resolution O 1s XPS spectrum can be divided into four peaks. The peaks at 530.3, 531.5, 532.2, and 533.13eV are typically attributed to metal–oxygen bonds, carbon–oxygen bonds, sulfur–oxygen bonds, and oxygen in hydroxyl groups, respectively [13]. The peaks at 168.7 eV and 169.9 eV in the S 2p spectrum indicate the presence of SO42− anions in LDH (Figure 4f) [17]. SO42 displays a larger size when compared with other intercalated anions (Cl, CO32−, and NO3), which may result in increased spacing between the hydroxide layers, thus increasing the electrochemically active surface area of the catalyst, as well as its catalytic property [6,16]. Moreover, SO42− processes the higher negative charge and the higher electron-donating ability, which might also help facilitate the OER activity of the corresponding LDH materials [17].
The OER catalytic performance of ZIF-9(III), CoFe LDH, and commercial RuO2 catalysts were studied in a standard three-electrode setup using a 1.0 M KOH solution as an electrolyte. As seen in Figure 5a, CoFe LDH nanosheets exhibit a lower overpotential of 218 mV at 10 mA cm−2 than those of ZIF-9(III) (373 mV) and commercial RuO2 (285 mV). It is widely accepted that Fe3+ ions play an essential role in improving the electrochemical performance of Co LDH OER catalysts. For example, the strong electronic interactions between the Co and Fe elements have been demonstrated from the fact that Fe incorporation correlates with an anodic shift in the Co2+/Co3+ redox wave [8]. In addition, Fe3+ can stabilize the LDH structure due to the similar ionic radii of the Co and Fe ions in the hydroxide layer [40]. Thus, the enhanced catalytic performance can be attributed to the hierarchical structure composed of ultrathin nanosheets and the synergistic effect of Co and Fe [41]. In addition, the low value of the Tafel slope indicates that the rate-determining step is the final part of the multiple electron reactions, which is commonly a sign of great OER activity for catalysts [4]. As can be seen from Figure 5b, the slope of the Tafel plot of the CoFe LDH (59.9 mV dec1) is lower than ZIF-9 (69.1 mV dec−1), and commercial RuO2 (85.8 mV dec−1). Correspondingly, the Tafel slope of CoFe LDH suggests that the rate-determining step is the third electron transfer step, a step-generating MOOH intermediate [42]. Significantly, CoFe LDH performs competitively when compared with most of the catalysts reported recently (Figure 5c and Table S1).
To further investigate charge-transport kinetics, the electrochemical impedance spectroscopy (EIS) of electrodes was tested. The Nyquist plots of as-synthetized electrodes and the relevant equivalent circuit model in the form of Rs, (Q1, Ro) and (Q2, Rct) are displayed in Figure 5d. Rs denotes the solution resistance, and Ro denotes the oxide film resistance, which elaborate the electron transport of the electrode material/catalyst interface and the catalyst/electrolyte interface, respectively [43]. The charge-transfer impedance (Rct) indicates the charge-transfer kinetics of the OER catalytic process and further evaluates the OER activity of the electrocatalyst. The CoFe LDH exhibits a smaller Rct value (1.3 Ω cm2) than ZIF-9(III) (7.94 Ω cm2), indicating the low mass-transfer resistance and fast charge-transfer rate for the CoFe LDH. The electrochemical active surface areas (ECSA) of the as-prepared samples were evaluated by calculating the double-layer capacitance (Cdl), according to cyclic voltammograms (CV) at different scan rates (Figure S5). As seen from Figure 5e, CoFe LDH has a higher Cdl of 26.3 mF cm−2, which is superior to ZIF-9(III) (5.6 mF cm−2), suggesting that more exposed active sites in the CoFe LDH arise. Indeed, the turnover frequency (TOF) is an important parameter for catalysts. Generally, there are two approaches to determine the TOF, denoted by TOF total metal (TOFtm) and TOFredox [44,45]. As displayed in Figure S7, the TOF change trends for CoFe LDH and ZIF-9(III) catalysts are positively correlated with overpotential. Moreover, the TOF value of CoFe LDH is much higher than that of ZIF-9(III) at the same overpotential, confirming its superior intrinsic activity. Specifically, CoFe LDH nanosheets exhibit a higher TOF value, of 0.2716 s–1 at an overpotential of 300 mV, than that of ZIF-9(III) (0.0009 s–1). The molar number (n) is calculated based on the inductively coupled plasma-mass spectrometry (ICP-MS) results [13,46]. For CoFe LDH, the content of Co and Fe is 37.23 wt. % and 27.60 wt. %, respectively. For ZIF-9(III), the content of Co and Fe is 39.98 wt. % and 2.05 wt. %, respectively.
The catalytic stability of samples is another crucial parameter for OER. In Figure 5f, the durability of the CoFe LDH electrode is observed using chronopotentiometric measurements at a current density of 10 mA cm−2 within 55 h. Therefore, the catalytic stability of CoFe LDH is demonstrated in the alkaline electrochemical conditions for the OER. In order to further justify the structural stability of hierarchical CoFe LDH, the XRD pattern and SEM image for the CoFe LDH were provided after the OER tests. As shown in Figure S8a, the structure of catalysts can be ascribed to that of LDH [9]. Meanwhile, the catalyst still maintains hierarchical nanosheets after the OER process (Figure S8b). The results suggest that the hierarchical structure can improve the structural stability of LDH.
The outstanding OER performance of CoFe LDH can be attributed to the following aspects. Firstly, the co-existence of Co and Fe ions in the host layers can provide abundant redox reactions during electrochemical processes. Secondly, the large interlayer distance created using intercalated SO42− anions can give the advantage of exposure of more active sites. Thirdly, the hierarchical structure can improve the structural stability of LDH.

3. Experimental Procedure and Methods

3.1. Materials

Benzimidazole (H-PhIM; 98%), cobalt nitrate hexahydrate (Co(NO3)2⋅6H2O; 99%), ferrous sulfate heptahydrate (FeSO4·7H2O, 99%), potassium hydroxide (KOH; 90%), and dehydrated ethanol (EtOH; 99.7%) were used from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China). Nafion solution (5 wt. %) was supplied by DuPont Co., Ltd. (Circleville, OH, USA). The above chemicals were of analytical grade and were used directly without further purification. Ruthenium (IV) oxide (RuO2) was synthesized via directly calcining Ruthenium(III) chloride anhydrous (RuCl3, Ru content 45–55%) in the air at 400 °C.

3.2. Synthesis of Compounds

3.2.1. Preparation of 2D ZIF-9(III)

Two-dimensional ZIF-9(III) was prepared using a facile precipitation method with slight modifications, according to our previous report [29]. Firstly, 4 mmol H-PhIM and 2 mmol Co (NO3)2⋅6H2O were both dissolved in 100 mL deionized (DI) water under constant stirring at ambient temperature. Then, 20 mL ammonia water was added into the above solution after 5 min, and the resulting solution was continuously stirred for 45 min. After the solution was left to stand for 6 h, the solid powder was collected using centrifugation, washed with ethanol three times, and finally vacuum dried at 60 °C.

3.2.2. Synthesis of Hierarchical CoFe LDH Nanosheets

Firstly, the obtained 80 mg 2D ZIF-9(III) precursor was dispersed in a mixed solution of ethanol and DI water (the volume ratio of ethanol and water was 9:1). Then, the 50 mg FeSO4·7H2O was added into the above solution and stirred for 1 h. Finally, the resulting brown powder was washed three times with ethanol and dried in an oven at 60 °C overnight.

3.3. Physicochemical Characterization

The powder X-ray diffraction (XRD) data of as-prepared samples were collected from a Rigaku D/Max-3c X-ray diffractometer with Cu Kα radiation (λ = 1.5406 Å) at a range between 5 and 70° at a scan step of 5°/min. The morphologies and elemental compositions were obtained using field emission scanning electron microscopy (FE-SEM; SU8220) and Transmission Electron Microscopy (TEM; Model JEM-2011, JEOL) equipped with a Rontec energy-dispersive X-ray (EDX) system. The X-ray photoelectron spectrometer (XPS; ESCA-LAB250) with Al Kα radiation was used to characterize the chemical states (elementary composition and the valence state characteristics) of the catalysts, and binding energy corrections on high-resolution scans were calibrated by referencing the C 1s peak (284.8 eV). The functional groups of as-synthesized samples were analyzed at room temperature using Fourier transform infrared spectroscopy (FTIR; PerkinElmer) using KBr as the reference sample. Raman spectra of catalysts were obtained using inVia Quotation Evolution (Renishaw) equipment with laser source of 514 nm. The thickness of the as-fabricated nanosheets was determined using atomic force microscopy (AFM; Bruker). The N2 adsorption/desorption measurements were performed on a Micromeritics ASAP 2020 analyzer (GA, USA) at the liquid nitrogen temperature (77 K), after degassing the samples for 12 h under vacuum at 120 °C. The specific surface areas and pore size analysis of the samples were calculated from the N2-sorption isotherms via the Brunauer-Emmett-Teller (BET) and Barrett–Joyner–Halenda (BJH) method, respectively. The Co and Fe contents were tested using inductively coupled plasma-mass spectrometry (ICP-MS).

3.4. Electrochemical Measurements

All electrochemical properties of OER were demonstrated in 1.0 M KOH electrolyte (pH = 13.6) in a conventional three-electrode system at room temperature using a CHI760E electrochemical workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China). Graphite rod and Ag/AgCl with 1.0 M KOH (0.1989 V vs. RHE) were used as the counter electrode and the reference electrode, respectively. The carbon paper (CP) electrode (surface = 0.5 cm2) modified with catalysts was used as the working electrode. Prior to the test, the CP electrode was cleaned in a solution of diluted hydrochloric acid, followed by rinsing with deionized water. To obtain the uniformly distributed catalyst ink solution, catalyst (2 mg) and Nafion solution (25 μL, 5 wt. %) were dispersed in 250 μL mixture solution of DI water and absolute ethanol (1:1). Homogeneous ink was obtained using 30 min of ultrasonication, dropping on the surface of a carbon paper electrode with a load of 0.3 mg cm−2, and then drying at room temperature.
The measured potentials were converted according to the equation: Evs. RHE = (Evs. Ag/AgCl + 0.059 × pH + 0.1989) V. The overpotential was calculated as: η = ERHE − 1.23 V. Tafel slopes were derived from the linear sweep voltammetry (LSV) curves and computed as η = blogj + a, while the η, b, and j represented overpotential, Tafel slope, and current density, respectively. The LSV polarization curves were conducted at a scan rate of 5 mV/s from 0.2 V to 0.8 V (vs. Ag/AgCl) with 95% ohmic potential drop (iR) correction. The OER stability was tested using chronopotentiometry (CP) at a constant current density of 10 mA cm−2 for 55 h without iR-compensation. Electrochemical impedance spectroscopy (EIS) was carried out over a frequency range of 105–10−2 Hz at potential 1.52 V (vs. RHE) for CoFe LDH and 1.62 V (vs. RHE) for ZIF-9(III). The electrochemical double-layer capacitance (Cdl) curves of catalysts were measured using cyclic voltammetry in a non-Faradaic region (1.10-1.20 V vs. RHE) at varied scan rates of 10, 20, 30, 40, 50, and 60 mV s−1 based on the following equation: Cdl =(jajc)/(2 × v), where ja, jc, and v correspond to the current density of anode, cathode, and scan rate, respectively. The TOF was calculated using the following equation: TOF = jA/(4nF), where j (A cm−2) is current density, A is surface area (0.5 cm2), F is the Faraday constant of 96,485 C mol−1, and n is the molar number of total metal ions.

4. Conclusions

To summarize, we developed a facile etching–coprecipitation method to drive the conversion of 2D ZIF-9(III) into sulfate-intercalated hierarchical CoFe LDH nanosheets. The CoFe LDH exhibits excellent electrocatalytic performance and a high stability for OER in 1.0 M KOH. Specifically, the CoFe LDH can achieve a current density of 10 mA cm−2 at a low overpotential of 218 mV with a small Tafel slope of 59.9 mV dec1. The excellent electrocatalytic performance of CoFe LDH can be mainly attributed to the hierarchical structure, the synergistic effect between Co and Fe ions, and the large interlayer space distance originating from SO42− ions. The current work provides a novel method for the rational design of LDH catalysts with desirable nanostructure and improved activity. Moreover, the CoFe LDH catalyst can be incorporated in an industrially relevant scale-type electrolyzer due to the high electrocatalytic activity, stability, and low cost.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12070688/s1, Figure S1: FE-SEM images in different magnifications of ZIF-9(III); Figure S2: High magnification FE-SEM images of CoFe LDH; Figure S3: AFM topography showing several nanosheets with a thickness (a,c) ZIF-9, (b,d) LDH; Figure S4: HRTEM image of CoFe LDH nanosheets; Figure S5: (a) N2 adsorption-desorption isotherms and (b) the correspondin pore size distributions derived from the desorption isotherms of the CoFe LDH sample; Figure S6: Cyclic voltammograms (CVs) curves of ZIF-9(III) and CoFe LDH at scan rate of 10, 20, 30, 40, 50 and 60 mV s−1; Figure S7: The turnover frequency (TOF) versus of CoFe LDH and ZIF-9(III); Table S1: Comparison study of CoFe LDH in this work and previously reported similar materials toward oxygen evolution reaction. References [9,10,11,12,13,27,38] are citated in the Supplementary Materials.

Author Contributions

G.X.: Investigation, data curation, writing—original draft. W.C.: Methodology, formal analysis, writing—original draft. Y.C.: Formal analysis. S.Z. Methodology. D.W.: Validation. D.C.: Conceptualization, resources, visualization, funding acquisition, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China [No. 22068008, 21661008], Guangxi Province Natural Science Foundation [No. 2020GXNSFAA159104, 2019GXNSFGA245003].

Data Availability Statement

Data available in a publicly accessible repository that does not issue DOIs or on request from the corresponding author.

Acknowledgments

We thanks the funding support from the Science Funding Committee of China and the Science and Technology Committee of Guangxi Province.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, D.; Li, P.; Lin, X.; Mckinley, A.; Kuang, Y.; Liu, W.; Lin, W.-F.; Sun, X.; Duan, X. Layered double hydroxide-based electrocatalysts for the oxygen evolution reaction: Identification and tailoring of active sites, and superaerophobic nanoarray electrode assembly. Chem. Soc. Rev. 2021, 50, 8790–8817. [Google Scholar] [CrossRef] [PubMed]
  2. Lv, L.; Yang, Z.; Chen, K.; Wang, C.; Xiong, Y. 2D Layered Double Hydroxides for Oxygen Evolution Reaction: From Fundamental Design to Application. Adv. Energy Mater. 2019, 9, 1803358. [Google Scholar] [CrossRef]
  3. Li, J.; Jiang, S.; Shao, M.; Wei, M. Host-Guest Engineering of Layered Double Hydroxides towards Efficient Oxygen Evolution Reaction: Recent Advances and Perspectives. Catalysts 2018, 8, 214. [Google Scholar] [CrossRef] [Green Version]
  4. Yu, J.; Yu, F.; Yuen, M.-F.; Wang, C. Two-dimensional layered double hydroxides as a platform for electrocatalytic oxygen evolution. J. Mater. Chem. A 2021, 9, 9389–9430. [Google Scholar] [CrossRef]
  5. Wang, J.; Kong, H.; Zhong, H.; Jiang, Y.; Guo, F.; Alonso-Vante, N.; Feng, Y. Recent Progress on Transition Metal Based Layered Double Hydroxides Tailored for Oxygen Electrode Reactions. Catalysts 2021, 11, 1394. [Google Scholar] [CrossRef]
  6. Karmakar, A.; Karthick, K.; Sankar, S.S.; Kumaravel, S.; Madhu, R.; Kundu, S. A vast exploration of improvising synthetic strategies for enhancing the OER kinetics of LDH structures: A review. J. Mater. Chem. A 2021, 9, 1314–1352. [Google Scholar] [CrossRef]
  7. Zhou, Y.; Hu, J.; Yang, L.; Gao, Q. Recent advances of two-dimensional CoFe layered-double-hydroxides for electrocatalytic water oxidation. Chin. Chem. Lett. 2022, 33, 2845–2855. [Google Scholar] [CrossRef]
  8. Burke, M.S.; Kast, M.G.; Trotochaud, L.; Smith, A.M.; Boettcher, S.W. Cobalt–Iron (Oxy)hydroxide Oxygen Evolution Electrocatalysts: The Role of Structure and Composition on Activity, Stability, and Mechanism. J. Am. Chem. Soc. 2015, 137, 3638–3648. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, R.; Zhou, Y.; Xing, Y.; Li, D.; Jiang, D.; Chen, M.; Shi, W.; Yuan, S. Synergistic coupling of CoFe-LDH arrays with NiFe-LDH nanosheet for highly efficient overall water splitting in alkaline media. Appl. Catal. B 2019, 253, 131–139. [Google Scholar] [CrossRef]
  10. Wang, Z.; Liu, W.; Hu, Y.; Guan, M.; Xu, L.; Li, H.; Bao, J.; Li, H. Cr-doped CoFe layered double hydroxides: Highly efficient and robust bifunctional electrocatalyst for the oxidation of water and urea. Appl. Catal. B 2020, 272, 118959. [Google Scholar] [CrossRef]
  11. Wang, Y.; Zhang, Y.; Liu, Z.; Xie, C.; Feng, S.; Liu, D.; Shao, M.; Wang, S. Layered Double Hydroxide Nanosheets with Multiple Vacancies Obtained by Dry Exfoliation as Highly Efficient Oxygen Evolution Electrocatalysts. Angew. Chem. Int. Ed. 2017, 56, 5867–5871. [Google Scholar] [CrossRef]
  12. Feng, L.; Li, A.; Li, Y.; Liu, J.; Wang, L.; Huang, L.; Wang, Y.; Ge, X. A Highly Active CoFe Layered Double Hydroxide for Water Splitting. ChemPlusChem 2017, 82, 483–488. [Google Scholar] [CrossRef] [PubMed]
  13. Nie, F.; Li, Z.; Dai, X.; Yin, X.; Gan, Y.; Yang, Z.; Wu, B.; Ren, Z.; Cao, Y.; Song, W. Interfacial electronic modulation on heterostructured NiSe@CoFe LDH nanoarrays for enhancing oxygen evolution reaction and water splitting by facilitating the deprotonation of OH to O. Chem. Eng. J. 2022, 431, 134080. [Google Scholar] [CrossRef]
  14. Dang, L.; Liang, H.; Zhuo, J.; Lamb, B.K.; Sheng, H.; Yang, Y.; Jin, S. Direct Synthesis and Anion Exchange of Noncarbonate-Intercalated NiFe-Layered Double Hydroxides and the Influence on Electrocatalysis. Chem. Mater. 2018, 30, 4321–4330. [Google Scholar] [CrossRef]
  15. Hunter, B.M.; Hieringer, W.; Winkler, J.R.; Gray, H.B.; Müller, A.M. Effect of interlayer anions on [NiFe]-LDH nanosheet water oxidation activity. Energy Environ. Sci. 2016, 9, 1734–1743. [Google Scholar] [CrossRef] [Green Version]
  16. Xu, Y.; Hao, Y.; Zhang, G.; Lu, Z.; Han, S.; Li, Y.; Sun, X. Room-temperature synthetic NiFe layered double hydroxide with different anions intercalation as an excellent oxygen evolution catalyst. RSC Adv. 2015, 5, 55131–55135. [Google Scholar] [CrossRef]
  17. Jiang, X.-X.; Xue, J.-Y.; Zhao, Z.-Y.; Li, C.; Li, F.-L.; Cao, C.; Niu, Z.; Gu, H.-W.; Lang, J.-P. Ultrathin sulfate-intercalated NiFe-layered double hydroxide nanosheets for efficient electrocatalytic oxygen evolution. RSC Adv. 2020, 10, 12145–12150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Wu, H.; Wang, J.; Jin, W.; Wu, Z. Recent development of two-dimensional metal–organic framework derived electrocatalysts for hydrogen and oxygen electrocatalysis. Nanoscale 2020, 12, 18497–18522. [Google Scholar] [CrossRef]
  19. Xia, W.; Mahmood, A.; Zou, R.; Xu, Q. Metal–organic frameworks and their derived nanostructures for electrochemical energy storage and conversion. Energy Environ. Sci. 2015, 8, 1837–1866. [Google Scholar] [CrossRef]
  20. Wu, H.B.; Lou, X.W. Metal-organic frameworks and their derived materials for electrochemical energy storage and conversion: Promises and challenges. Sci. Adv. 2017, 3, eaap9252. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, J.; Li, F.; Chen, W.; Wang, C.; Cai, D. Facile synthesis of hollow Co3O4-embedded carbon/reduced graphene oxides nanocomposites for use as efficient electrocatalysts in oxygen evolution reaction. Electrochim. Acta 2019, 300, 123–130. [Google Scholar] [CrossRef]
  22. Wang, C.; Chen, W.; Yuan, D.; Qian, S.; Cai, D.; Jiang, J.; Zhang, S. Tailoring the nanostructure and electronic configuration of metal phosphides for efficient electrocatalytic oxygen evolution reactions. Nano Energy 2020, 69, 104453. [Google Scholar] [CrossRef]
  23. Liu, M.; Kong, L.; Wang, X.; He, J.; Bu, X.-H. Engineering Bimetal Synergistic Electrocatalysts Based on Metal–Organic Frameworks for Efficient Oxygen Evolution. Small 2019, 15, 1903410. [Google Scholar] [CrossRef] [PubMed]
  24. Rinawati, M.; Wang, Y.-X.; Chen, K.-Y.; Yeh, M.-H. Designing a spontaneously deriving NiFe-LDH from bimetallic MOF-74 as an electrocatalyst for oxygen evolution reaction in alkaline solution. Chem. Eng. J. 2021, 423, 130204. [Google Scholar] [CrossRef]
  25. Zhang, J.; Yu, L.; Chen, Y.; Lu, X.F.; Gao, S.; Lou, X.W. Designed Formation of Double-Shelled Ni–Fe Layered-Double-Hydroxide Nanocages for Efficient Oxygen Evolution Reaction. Adv. Mater. 2020, 32, 1906432. [Google Scholar] [CrossRef]
  26. Li, J.-G.; Sun, H.; Lv, L.; Li, Z.; Ao, X.; Xu, C.; Li, Y.; Wang, C. Metal–Organic Framework-Derived Hierarchical (Co,Ni)Se2@NiFe LDH Hollow Nanocages for Enhanced Oxygen Evolution. ACS Appl. Mater. Interfaces 2019, 11, 8106–8114. [Google Scholar] [CrossRef]
  27. Hu, L.; Xiao, R.; Wang, X.; Wang, X.; Wang, C.; Wen, J.; Gu, W.; Zhu, C. MXene-induced electronic optimization of metal-organic framework-derived CoFe LDH nanosheet arrays for efficient oxygen evolution. Appl. Catal. B 2021, 298, 120599. [Google Scholar] [CrossRef]
  28. Li, J.; Wang, L.; He, H.; Chen, Y.; Gao, Z.; Ma, N.; Wang, B.; Zheng, L.; Li, R.; Wei, Y.; et al. Interface construction of NiCo LDH/NiCoS based on the 2D ultrathin nanosheet towards oxygen evolution reaction. Nano Res. 2022, 1–10. [Google Scholar] [CrossRef]
  29. Chen, W.; Wang, C.; Su, S.; Wang, H.; Cai, D. Synthesis of ZIF-9(III)/Co LDH layered composite from ZIF-9(I) based on controllable phase transition for enhanced electrocatalytic oxygen evolution reaction. Chem. Eng. J. 2021, 414, 128784. [Google Scholar] [CrossRef]
  30. Zhao, P.; Lampronti, G.I.; Lloyd, G.O.; Wharmby, M.T.; Facq, S.; Cheetham, A.K.; Redfern, S. A T. Phase Transitions in Zeolitic Imidazolate Framework 7: The Importance of Framework Flexibility and Guest-Induced Instability. Chem. Mater. 2014, 26, 1767–1769. [Google Scholar] [CrossRef]
  31. Lin, Y.; Wang, H.; Peng, C.-K.; Bu, L.; Chiang, C.-L.; Tian, K.; Zhao, Y.; Zhao, J.; Lin, Y.-G.; Lee, J.-M.; et al. Co-Induced Electronic Optimization of Hierarchical NiFe LDH for Oxygen Evolution. Small 2020, 16, 2002426. [Google Scholar] [CrossRef]
  32. Ebrahimi, A.; Mansournia, M. Cost-effective fabrication of thermal- and chemical-stable ZIF-9 nanocrystals at ammonia atmosphere. J. Phys. Chem. Solids 2017, 111, 12–17. [Google Scholar] [CrossRef]
  33. Wang, H.; Jin, Z. Rational design W-doped Co-ZIF-9 based Co3S4 composite photocatalyst for efficient visible-light-driven photocatalytic H2 evolution. Sustain. Energy Fuels 2019, 3, 173–183. [Google Scholar] [CrossRef]
  34. Jia, H.; Zhao, S.; Shi, Y.; Zhu, K.; Gao, P.; Zhu, L. Mechanisms for light-driven evolution of environmentally persistent free radicals and photolytic degradation of PAHs on Fe(III)-montmorillonite surface. J. Hazard. Mater. 2019, 362, 92–98. [Google Scholar] [CrossRef] [PubMed]
  35. Carja, G.; Grosu, E.F.; Mureseanu, M.; Lutic, D. A family of solar light responsive photocatalysts obtained using Zn2+ Me3+ (Me = Al/Ga) LDHs doped with Ga2O3 and In2O3 and their derived mixed oxides: A case study of phenol/4-nitrophenol decomposition. Catal. Sci. Technol. 2017, 7, 5402–5412. [Google Scholar] [CrossRef]
  36. Fahami, A.; Beall, G.W. Mechanosynthesis and characterization of Hydrotalcite like Mg–Al–SO4-LDH. Mater. Lett. 2016, 165, 192–195. [Google Scholar] [CrossRef] [Green Version]
  37. Jayaramulu, K.; Masa, J.; Morales, D.M.; Tomanec, O.; Ranc, V.; Petr, M.; Wilde, P.; Chen, Y.-T.; Zboril, R.; Schuhmann, W.; et al. Ultrathin 2D Cobalt Zeolite-Imidazole Framework Nanosheets for Electrocatalytic Oxygen Evolution. Adv. Sci. 2018, 5, 1801029. [Google Scholar] [CrossRef] [Green Version]
  38. Zhu, K.; Chen, J.; Wang, W.; Liao, J.; Dong, J.; Chee, M.O.L.; Wang, N.; Dong, P.; Ajayan, P.M.; Gao, S.; et al. Etching-Doping Sedimentation Equilibrium Strategy: Accelerating Kinetics on Hollow Rh-Doped CoFe-Layered Double Hydroxides for Water Splitting. Adv. Funct. Mater. 2020, 30, 2003556. [Google Scholar] [CrossRef]
  39. Shi, Y.; Du, W.; Zhou, W.; Wang, C.; Lu, S.; Lu, S.; Zhang, B. Unveiling the Promotion of Surface-Adsorbed Chalcogenate on the Electrocatalytic Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2020, 59, 22470–22474. [Google Scholar] [CrossRef]
  40. Yang, F.; Sliozberg, K.; Sinev, I.; Antoni, H.; Bähr, A.; Ollegott, K.; Xia, W.; Masa, J.; Grünert, W.; Cuenya, B.R.; et al. Synergistic Effect of Cobalt and Iron in Layered Double Hydroxide Catalysts for the Oxygen Evolution Reaction. ChemSusChem 2017, 10, 156–165. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Yang, M.; Jiang, X.; Lu, W.; Xing, Y. Self-supported hierarchical CoFe-LDH/NiCo2O4/NF core-shell nanowire arrays as an effective electrocatalyst for oxygen evolution reaction. J. Alloys Compd. 2020, 818, 153345. [Google Scholar] [CrossRef]
  42. Liu, H.; Zha, M.; Liu, Z.; Tian, J.; Hu, G.; Feng, L. Synergistically boosting the oxygen evolution reaction of an Fe-MOF via Ni doping and fluorination. Chem. Commun. 2020, 56, 7889–7892. [Google Scholar] [CrossRef]
  43. Li, F.; Tian, Y.; Su, S.; Wang, C.; Li, D.-S.; Cai, D.; Zhang, S. Theoretical and experimental exploration of tri-metallic organic frameworks (t-MOFs) for efficient electrocatalytic oxygen evolution reaction. Appl. Catal. B 2021, 299, 120665. [Google Scholar] [CrossRef]
  44. Battiato, S.; Urso, M.; Cosentino, S.; Pellegrino, A.L.; Mirabella, S.; Terrasi, A. Optimization of Oxygen Evolution Reaction with Electroless Deposited Ni–P Catalytic Nanocoating. Nanomaterials 2021, 11, 3010. [Google Scholar] [CrossRef] [PubMed]
  45. Stevens, M.B.; Enman, L.J.; Batchellor, A.S.; Cosby, M.R.; Vise, A.E.; Trang, C.D.M.; Boettcher, S.W. Measurement Techniques for the Study of Thin Film Heterogeneous Water Oxidation Electrocatalysts. Chem. Mater. 2017, 29, 120–140. [Google Scholar] [CrossRef]
  46. Wang, C.-P.; Feng, Y.; Sun, H.; Wag, Y.; Yin, J.; Yao, Z.; Bu, X.-H.; Zhu, J. Self-Optimized Metal–Organic Framework Electrocatalysts with Structural Stability and High Current Tolerance for Water Oxidation. ACS Catal. 2021, 11, 7132–7143. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthetic procedure of CoFe LDH nanosheets.
Figure 1. Schematic illustration of the synthetic procedure of CoFe LDH nanosheets.
Catalysts 12 00688 g001
Figure 2. XRD patterns of (a) ZIF-9(III), simulated ZIF-9(III) and (b) CoFe LDH, (c) FTIR spectrum, and (d) Raman spectroscopy of ZIF-9(III) and CoFe LDH.
Figure 2. XRD patterns of (a) ZIF-9(III), simulated ZIF-9(III) and (b) CoFe LDH, (c) FTIR spectrum, and (d) Raman spectroscopy of ZIF-9(III) and CoFe LDH.
Catalysts 12 00688 g002
Figure 3. (a) FE-SEM image of ZIF-9(III), (bd) FE-SEM images, (e,f) TEM and (g) HRTEM images in different magnifications of CoFe LDH, and (h) EDX elemental mapping images for CoFe LDH.
Figure 3. (a) FE-SEM image of ZIF-9(III), (bd) FE-SEM images, (e,f) TEM and (g) HRTEM images in different magnifications of CoFe LDH, and (h) EDX elemental mapping images for CoFe LDH.
Catalysts 12 00688 g003
Figure 4. (a) XPS survey spectrum and (bf) high-resolution XPS spectra of the Co 2p, Fe 2p, C 1s, O 1s, and S 2p regions of the CoFe LDH sample.
Figure 4. (a) XPS survey spectrum and (bf) high-resolution XPS spectra of the Co 2p, Fe 2p, C 1s, O 1s, and S 2p regions of the CoFe LDH sample.
Catalysts 12 00688 g004
Figure 5. (a) LSV curves of ZIF-9(III), CoFe LDH, and commercial RuO2, (b) the corresponding Tafel plots, (c) overpotentials at 10 mA cm−2 and Tafel slopes of some previously reported CoFe LDH catalysts, (d) EIS curves and fitting results for CoFe LDH at 1.52V vs. RHE and ZIF-9(III) at 1.62V vs. RHE, (e) electrochemical double-layer capacitance (Cdl) of ZIF-9(III) and CoFe LDH, (f) chronopotentiometry of CoFe LDH.
Figure 5. (a) LSV curves of ZIF-9(III), CoFe LDH, and commercial RuO2, (b) the corresponding Tafel plots, (c) overpotentials at 10 mA cm−2 and Tafel slopes of some previously reported CoFe LDH catalysts, (d) EIS curves and fitting results for CoFe LDH at 1.52V vs. RHE and ZIF-9(III) at 1.62V vs. RHE, (e) electrochemical double-layer capacitance (Cdl) of ZIF-9(III) and CoFe LDH, (f) chronopotentiometry of CoFe LDH.
Catalysts 12 00688 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiao, G.; Chen, W.; Cai, Y.; Zhang, S.; Wang, D.; Cai, D. Facile Synthesis of Sulfate-Intercalated CoFe LDH Nanosheets Derived from Two-Dimensional ZIF-9(III) for Promoted Oxygen Evolution Reaction. Catalysts 2022, 12, 688. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12070688

AMA Style

Xiao G, Chen W, Cai Y, Zhang S, Wang D, Cai D. Facile Synthesis of Sulfate-Intercalated CoFe LDH Nanosheets Derived from Two-Dimensional ZIF-9(III) for Promoted Oxygen Evolution Reaction. Catalysts. 2022; 12(7):688. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12070688

Chicago/Turabian Style

Xiao, Guolei, Weibin Chen, Yaming Cai, Shifan Zhang, Di Wang, and Dandan Cai. 2022. "Facile Synthesis of Sulfate-Intercalated CoFe LDH Nanosheets Derived from Two-Dimensional ZIF-9(III) for Promoted Oxygen Evolution Reaction" Catalysts 12, no. 7: 688. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12070688

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

Article Metrics

Back to TopTop