Next Article in Journal
Investigation of the Mechanical Properties and Microstructure of the Co40NiCrMo Alloy Used for STACERs and Prepared by the CSPB Process and the Winding and Stabilization Method
Previous Article in Journal
Phase Field Study on the Spinodal Decomposition of β Phase in Zr–Nb-Ti Alloys
Previous Article in Special Issue
Electrocatalytic Properties of Ni(II) Schiff Base Complex Polymer Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heterostructure of NiFe@NiCr-LDH for Active and Durable Oxygen Evolution Reactions in Alkaline Media

1
New Energy Research Institute, School of Environment and Energy, South China University of Technology, Guangzhou Higher Education Mega Center, Guangzhou 510006, China
2
State Key Laboratory of Subtropical Building Science, South China University of Technology, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(8), 2968; https://doi.org/10.3390/ma16082968
Submission received: 8 March 2023 / Revised: 6 April 2023 / Accepted: 6 April 2023 / Published: 8 April 2023
(This article belongs to the Special Issue Advances in Electrocatalysts: Synthesis and Applications)

Abstract

:
Developing cost-effective, efficient, and durable catalysts for oxygen evolution reactions (OER) is the key for promoting large-scale H2 production through electrochemical water splitting. Herein, we report a facile method for fabricating an NiFe@NiCr-LDH catalyst toward alkaline OER. The electronic microscopy technique revealed that it has a well-defined heterostructure at the interface between the NiFe and NiCr phases. In 1.0 M KOH, the as-prepared NiFe@NiCr-LDH catalyst shows excellent catalytic performance, evidenced by an overpotential of 266 mV at the current density of 10 mA cm−2 and a small Tafel slope of 63 mV dec−1; both are comparable with the RuO2 benchmark catalyst. It also exhibits robust durability in long-term operation, manifested by a 10% current decay in 20 h, which is superior to that of the RuO2 catalyst. Such excellent performance is attributed to the interfacial electron transfer that occurs at the interfaces of the heterostructure, and the Fe(III) species facilitate the formation of Ni(III) species as active sites in NiFe@NiCr-LDH. This study offers a feasible strategy for preparing a transition metal-based LDH catalyst for OER toward H2 production and other electrochemical energy technologies.

Graphical Abstract

1. Introduction

Promoting green energy technologies with low-carbon or even carbon-free emission is essential to resolve the global warming challenge and ubiquitous environmental issues [1,2]. H2 has been widely regarded as one of the most promising green energy sources because it has an ultrahigh energy density and does not produce any carbon emission when using it [3,4]. Electrochemical water splitting has been proven to be one effective strategy for producing high-quality H2, yet developing efficient and durable catalysts for two half reactions including hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) is critical [5,6,7,8]. Nevertheless, compared to HER, the sluggish kinetics of OER, which has a more compldata nex four-electron transfer process, are the bottleneck that determines the water electrolysis efficiency [9,10]. Currently, RuO2 and IrO2 are the benchmark catalysts for OER, but the low abundance in Earth’s crust has significantly restricted their large-scale commercial application [11,12,13,14]. Moreover, both RuO2 and IrO2 are suffering from other issues that are detrimental to the practical application, e.g., RuO2 can form RuO42− and becomes dissolved under a high positive potential in the OER process, thus resulting in poor durability [15,16]. To that end, developing low-cost, efficient, and durable OER catalysts is imperative for realizing practical water splitting toward H2 generation.
Recently, transition metal layered double hydroxides (LDHs, whose chemical formula is denoted as [M(II)1−xM(III)x(OH)2][An−]x/n·yH2O, where M(II) can be Ni2+, Co2+, Zn2+, etc., M(III) can be Fe3+, Cr3+, V3+, etc., and An− is the charge balancing anion) have been attracting considerable interest as promising OER catalysts, thanks to their relatively low cost, their facile synthesis, and, more importantly, the well-defined two-dimensional structure that induces abundant active sites that can be exposed for OER [17,18,19,20]. Among these, NiFe-LDHs are the most widely explored, not only because of the high abundance of the Ni/Fe elements but also due to the fact that numerous NiOOH active species can be generated at the edge of NiFe-LDHs during the OER process [21,22,23]. To enhance the OER performance of NiFe-LDHs, various strategies have been documented, including morphology control, doping, defect engineering, intercalation, exfoliation, and so on [24,25,26,27,28,29]. A previous study revealed that introducing a third metal can lead to the charge redistribution and hence may enhance the OER performance [30,31,32]. For instance, Zhang et al. reported a catalyst of single Au atoms supported on NiFe-LDH (denoted as sAu/NiFe LDH), which had a small overpotential of 0.21 V at the current density of 10 mA cm−2 in 1 M KOH for OER; DFT calculation demonstrated that the Au single atoms can induce the formation of an active NiFeOOH layer on the LDHs’ edges, leading to improved adsorption energies of the OER intermediates on the Fe active sites [33]. In another study, Feng’s group prepared a ruthenium catalyst anchored on the surface of CoFe-LDHs (denoted as Ru/CoFe-LDHs), and it showed an overpotential of 198 mV at 10 mA cm−2 and a small Tafel slope of 39 mV dec−1 in 1.0 M KOH for OER; Operando X-ray absorption spectroscopy (XAS) disclosed that Ru maintained the oxidation state of IV in the OER process due to the optimal synergistic electron coupling, and such highly valent Ru species served as the active center [34].
Despite the above progress, the OER performance of NiFe-LDHs is still restricted by its physiochemical properties—particularly, the insulation properties of nickel hydroxides limit the charge transfer capability [22,27,35]. Therefore, the appropriate adjustment of the electronic structure of nickel hydroxides to improve the electric conductivity holds great potential for enhancing the electrocatalytic performance [36,37,38]. Cr(III) cations occupying the octahedral center (NiFe-LDHs have numerous such centers) present a unique electronic configuration (t2g3eg0) which can accelerate the charge transfer and hence can be beneficial for the OER process [38,39,40,41,42,43]. For example, Jin’s team synthesized a stable trimetallic NiFeCr-LDHs (molar ratio Ni:Fe:Cr = 6:2:1) catalyst on conductive carbon paper, and it demonstrated superior OER performance to that of NiFe-LDHs, with a low overpotential of 225 mV at a current density of 25 mA cm−2, thanks to the synergistic interaction in the metal center [44]. In another study, the Chen group developed NiCr-LDH as a bifunctional electrocatalyst for overall water splitting, and the optimized sample of Ni2Cr1-LDH showed extraordinary OER activity, evidenced by an ultralow overpotential of 319 mV at 100 mA cm−2 and outstanding durability at 1.55 V (vs. RHE) for 30 h in 1 M KOH; DFT calculations disclosed that the Cr3+ ions within the LDH layer served as charge transfer sites and thus boosted the intrinsic electrochemical activity [45].
Inspired by the above findings, herein, we report Ni-Fe and Ni-Cr coupled LDHs with a heterostructure (NiFe@NiCr-LDH) through a facile two-step hydrothermal approach as a high-efficiency OER catalyst. A series of characterizations revealed that a well-defined heterostructure is formed between the NiFe-LDH and NiCr-LDH phases in NiFe@NiCr-LDH. NiFe@NiCr-LDH exhibited excellent OER performance in 1 M KOH, manifested by a low overpotential of 266 mV at 10 mA cm−2 and a small Tafel slope of 63 mV dec−1, which are both comparable with those of the RuO2 benchmark catalyst, as well as superior long-term stability compared to that of RuO2. The XPS analysis before and after the OER test disclosed that the outer shell Fe(III) species facilitate the oxidation of Ni(II) into Ni(III) as active sites for OER, which cannot be achieved with NiCr@NiFe-LDH and other control samples.

2. Experimental Section

2.1. Materials

Ruthenium dioxide (RuO2) and 5% Nafion perfluorinated resin were bought from the Aladdin chemical company. Nickel (II) nitrate hexahydrate (Ni(NO3)2·6H2O, 99%), iron (III) nitrate nonahydrate (Fe(NO3)3·9H2O, 99%), chromium (III) nitrate nonahydrate (Cr(NO3)3·9H2O, 99%), sodium fluoride (NaF, ≥98.0%), and ethanol (99.0%) were purchased from Energy Chemical Company, China. Urea (CO(NH2)2, ≥99%) and ammonium fluoride (NH4F, >99%) were purchased from Macklin Chemical Company. All solutions were prepared using deionized water (DIW, the resistivity is 18.2 MΩ cm−1 at room temperature). All the chemicals were used as received, without further purification.

2.2. Synthesis of NiCr-LDH

NiCr-LDH was synthesized by following a modified hydrothermal approach [30,45,46]. Typically, 0.8 mL of Ni(NO3)2 solution (1.0 M) and 0.8 mL of Cr(NO3)3 solution (0.5 M) were mixed together into 3.4 mL DIW to form Solution A. Meanwhile, 5 mmol urea, 1 mmol NaF, and 2 mmol NH4F were added into 10 mL DIW under mild stirring for 5 min to prepare Solution B. Solutions A and B were then mixed and transferred into a 25 mL Teflon autoclave. Then, the liner was heated at 120 °C for 6 h, during which NiCr-LDH nanospheres slowly precipitated out. After that, the mixture was centrifuged and washed by DIW and ethanol, respectively, several times. After being dried overnight in vacuum, wathet blue powder as the final product was collected and denoted as NiCr-LDH.

2.3. Synthesis of NiFe@NiCr-LDH

On the basis of forming NiCr-LDH, NiFe@NiCr-LDH was further synthesized by a second hydrothermal step similar to the first one (See Scheme 1). Such sequential hydrothermal step is important for constructing a heterostructure between NiFe-LDH and NiCr-LDH instead of forming homogeneous layered triple hydroxide (LTH). Typically, 0.3 mL Ni(NO3)2 solution (1.0 M) and 0.2 mL Fe(NO3)3 solution (0.5 M) were added into 4.5 mL DIW to form Solution C. Meanwhile, 5 mmol urea and 3 mmol NH4F were added into another 10 mL of DIW under mild stirring for 5 min to prepare Solution D. Solutions C and D were then mixed and transferred into a 25 mL Teflon autoclave. After that, 60 mg NiCr-LDH powder was added into the autoclave with vigorous ultrasonic treatment for 5 min. Subsequently, the autoclave was heated at 120 °C for 2 h, during which NiFe-LDH nanosheets were precipitated on the surface of NiCr-LDH nanospheres. After that, the mixture was centrifuged, washed, and dried. The final product was collected and also denoted as NiFe@NiCr-LDH.

2.4. Characterizations

The morphologies of the samples were observed by a field-emission scanning electronic microscope (SEM, Tescan MIRA LMS), a field-emission transmission electronic microscope (TEM, Talos F200X operating at 200 kV), and high-angle annular dark-field scanning TEM (HAADF-STEM operating at 200 kV), and the elemental mapping was analyzed by energy dispersive spectroscopy (EDS, Tencai G2 F20 X-Twin). The mass content of the Ni, Cr, Fe, and O elements was determined by inductive coupled plasma optical emission spectroscopy (ICP-OES, Agilent 720ES). The physiochemical properties of the samples were characterized by high-resolution TEM (HRTEM), selected area electron diffraction (SAED, Talos F200X), Fourier transform infrared spectroscopy (FTIR, Thermo Nicolet iS 10), X-Ray Diffraction (XRD, Rigaku SmartLab using Cu Kα radiation at 10° min−1 in the 2θ range of 5°–90°), and X-ray photoelectron spectroscopy (XPS, PHI 5000 Versaprobe III).

2.5. Electrochemical Tests

All the electrochemical measurements were conducted on a CHI760D electrochemical workstation (Chenhua Instruments Co. Ltd., China) with a three-electrode standard system in 1.0 M KOH aqueous solution. The catalysts cast on a glassy carbon electrode (GCE) was adopted as the working electrode, which was prepared as follows: 5 mg of each sample was first dispersed in 1 mL solution made of 0.04 mL of Nafion, 0.24 mL of water, and 0.72 mL of ethanol. Then, the mixture was sonicated for 30 min to obtain the uniform suspension. Subsequently, 5 μL of the dispersion was dropwise cast onto GCE and dried to make a working electrode. A carbon rod and an Hg/HgO electrode were used as the counter electrode and the reference electrode, respectively. All the potentials were calibrated to a reversible hydrogen electrode (RHE) according to the following equation:
ERHE = EHg/HgO + (0.098 + 0.05916 × pH) V
where E is the potential versus RHE or Hg/HgO electrode.
Linear sweep voltammetry (LSV) polarization curves were recorded at a scan rate of 5 mV s−1 with iR compensation. The Tafel slope was calculated based on the following equation:
η = b log j + a
where η is the overpotential, b is the Tafel slope, a is the Tafel intercept, and j is the current density (mA cm−2). Electrochemical impedance spectroscopy (EIS) measurements were performed at an overpotential of 50 mV from 100 MHz to 1 Hz with an alternating current amplitude of 5 mV. The electrochemical double-layer capacitance (Cdl) was determined from the CV curves and measured in a potential range without Faradaic current by Cdl = Δj/Δv, where v is the scan rate (20~100 mV s−1) and Δj equals half of the difference between the anode and cathode current density (Δj = (ja − jc)/2). The long-term durability of the catalyst was investigated by chronoamperometry.

3. Results and Discussion

3.1. Characterization of the Samples

The NiFe@NiCr-LDH catalyst was first prepared by following the synthetic route in Scheme 1. Briefly, Cr(NO3)3, Ni(NO3)2, urea, and NH4F were reacted together through a hydrothermal approach to form NiCr-LDH first; then, NiFe-LDH was in situ formed onto it to yield NiFe@NiCr-LDH. In addition, a series of control samples including NiCr-LDH, NiFe-LDH, NiCrFe-LTH, and NiCr@NiFe-LDH were also prepared (Scheme S1). The detailed procedure can be found in the Supplementary Information. The mass loadings of Ni, Cr, and Fe were then determined by ICP-OES to be 35.1 wt.%, 11.0 wt.%, and 4.4 wt.%, respectively (Table S1).
The XRD patterns of the samples were then probed to obtain the crystal structure information, where the Ni2+, Fe3+, and/or Cr3+ ions occupy the octahedral center of the unit in the LDHs (Scheme 1b-c, Scheme S1c–e). As shown in Figure 1a, all the samples possess the peaks at 11.4°, 23.0°, 33.5°, 34.4°, 39.0°, 59.9°, and 61.2°, which can be attributed to the (003), (006), (101), (012), (015), (110), and (113) crystal planes of the LDHs (JCPDS no. 40-0215), respectively [30]. Moreover, the intensity of the peaks in NiFe-LDH is much higher than that of NiCr-LDH, while the FWHM values of the peaks in NiFe-LDH are lower than those of the other samples, even though the different LDH phases have a similar structure. It suggests that, the introduction of Cr(III) can significantly reduce the LDH’s crystallinity [47]. In addition, the peak of hydrated chromium hydroxide (Cr(OH)3·xH2O) located at 18.5° was detected in all the Cr(III)-containing LDH samples [48]. The hydroxide may partly dissolve in base media and form pores in LDHs, which is probably beneficial for exposing active sites for OER. FTIR spectroscopy was then performed to detect the interlayer ions and molecules. As shown in Figure 1b, the broad strong absorption band from 3700 cm−1 to 2500 cm−1 can be attributed to the stretching vibration of the O-H group. The peak at ~1651 cm−1 is assigned to the bending vibration of interlayer bound water, indicating the existence of hydrogen bonds. Other interlayer anions may include CO32− or HCO3−, and the corresponding sharp stretching vibration absorption peaks are located at ~1384 cm−1. The wide peak located at ~794 cm−1 demonstrates the existence of F [44]. Therefore, CO32−, HCO3, and F- are the main interlayer anions of all the LDH samples. The absorption peaks below 800 cm−1 are related to the bending and stretching modes of M-OH and M-OM in the brucite-like layers (M = Ni, Cr, or Fe); yet, the peak intensity can distinguish the NiFe-LDH phase and NiCr-LDH phase in the samples [42]. Nevertheless, NiCr@NiFe-LDH mainly inherits the feature from NiFe-LDH, while NiFe@NiCr-LDH preserves the characteristics from NiCr@LDH.
The morphology of the samples was then examined by the SEM technique. From low-magnification SEM images of NiFe-LDH and NiCr-LDH (Figure 1c and Figure S1), it can be noted that both NiFe-LDH and NiCr-LDH show a nanosheet primary structure, spherical secondary structure, and coral-like tertiary structure [49]. One distinction between them is that NiFe-LDH consists of vivid nanosheets, while NiCr-LDH has vague ones because of its lower crystallinity [45]. Another difference is the size of these nanospheres: the NiFe-LDH nanospheres’ diameter is ~2 μm, but the NiCr-LDH nanospheres’ diameter is smaller than 500 nm. Further, the other three samples have a distinctive morphology and vivid phase structure. The SEM images of the NiFe@NiCr-LDH nanospheres (Figure 1d–f) exhibit a combined morphology feature: relatively small NiFe-LDH nanosheets grown on the surface of NiCr-LDH spheres (1 μm); hence, NiFe@NiCr-LDH adopts the structural feature of both NiFe-LDH and NiCr-LDH. In comparison, NiCr@NiFe-LDH (Figure S2a–c) has a vague primary structure covered by external NiCr-LDH phases (Figure S1d–f) compared to that of NiFe-LDH (Figure S1a–c); NiCrFe-LTH (Figure S2d–f) has an NiFe-LDH-like morphology with a homogeneous structure that is quite different from heterogeneous NiFe@NiCr-LDH or NiCr@NiFe-LDH [50].
The detailed microstructure of the samples was then scrutinized by TEM and high-resolution TEM (HR-TEM) techniques. Figure 2a shows that the NiFe@NiCr-LDH nanosheets are stacked together, while similar stacking phenomena can also be observed in NiCr@NiFe-LDH and NiCrFe-LTH (Figure S3a–c) [30]. Further, the HR-TEM analysis can clearly differentiate the different phases (Figure S4e), as they possess different crystallinity [22]. As illustrated in Figure 2b,d, the NiCr-LDH phase (with vague lattice fringes in all places) and the NiFe-LDH phase (with well-defined lattice fringes everywhere) have identifiable boundaries, suggesting that a heterostructure is formed in both NiFe@NiCr-LDH and NiCr@NiFe-LDH. Interestingly, the lattice spacing of the NiFe-LDH (012) crystal phase is identical with that of the NiCr-LDH (012) phase. In contrast, NiCrFe-LTH displays homogeneous crystallinity (Figure S3d), reflecting its homogeneous structure [47]. In addition, the SAED of NiFe@NiCr-LDH (Figure 2c) exhibits the combined features, where the diffraction rings are from the crystal planes of small-sized NiCr-LDH phases and the diffraction spots are mainly attributed to the large and integral NiFe-LDH’s crystal phases, as observed in the SEM images and an HR-TEM image (Figure S4e) [42]. Therefore, NiFe@NiCr-LDH consists of primary NiCr-LDH phases and secondary NiFe-LDH phases. Instead, NiFe-LDH phases are dominant in NiCr@NiFe-LDH (Figure S4a), and the presence of more vague diffraction rings in NiCrFe-LTH’s pattern (Figure S4b) is due to the fact that Cr(III) is ubiquitous in the whole sample. Furthermore, such structural pattern in NiFe@NiCr-LDH is verified by the HAADF-STEM image and elemental mapping analysis (Figure 2e–i) [48]. One may notice that the Fe element has a somehow wider distribution region than the Cr element, indicating that NiFe@NiCr-LDH has more external Cr and internal Fe elements. In contrast, although the mass loading of Fe is higher than that of Cr in NiCr@NiFe-LDH, more Cr elements are observed in this sample, suggesting that it is in the outer layer (Figure S5), while the Ni, Cr, and Fe elements are homogeneously distributed in the NiCrFe-LTH sample (Figure S6). More importantly, the EDS line-scanning profile covers a whole NiFe@NiCr-LDH nanosphere (Figure S4c,d), and the elemental distribution can be clearly identified. As depicted in Figure S4d, both Ni and Cr have a peak shape distribution, but the intensity of Ni is much higher than that of Cr, suggesting that Ni is everywhere in the whole sample but Cr is mainly located in the center, and, more importantly, Fe has a slightly higher intensity at the edges than in the middle, suggesting it has more content in the outer layer.
Subsequently, the chemical composition and valence state of the samples were investigated by X-ray photoelectron spectroscopy (XPS). The existence of the Ni, Cr, Fe, C, O, and F elements can be proved distinctly in survey-scan spectra (Figure S7a). Further, more detailed structural information can be obtained through the high-resolution XPS spectra. In the core-level O 1s spectra (Figure 3a), the metal hydroxide (M-OH) species is identified at the peak binding energy of 531.4 eV, and interlayer bound water (H-OH) and metal oxide (M-O) species are also observed [51]. Nevertheless, the binding energy of the O 1s electrons exhibited no significant difference in NiFe-LDH, NiCr-LDH, and NiFe@NiCr-LDH, indicating that the M-OH species is dominant in the three samples. In the high-resolution C 1s (Figure S7b) and F 1s (Figure S7c) spectra of each sample, the presence of O-C=O (288.9 eV) and F (684.8 eV) can be readily identified in good echo with the findings in the XRD measurement, further attesting to the presence of these interlayer anions [44].
The high-resolution Ni2+ 2p spectra of the NiCr-LDH, NiFe-LDH, and NiFe@NiCr-LDH are presented in Figure 3b [52]. The binding energies of the Ni2+ 2p3/2 and 2p1/2 electrons in NiFe@NiCr-LDH are 855.7 eV and 873.0 eV, respectively, which are slightly higher than that in NiCr-LDH and lower than that in NiFe-LDH. It suggests that there is more electron transfer from Ni(II) to Fe(III) than from Ni(II) to Cr(III) because of iron’s higher electronegativity compared to chromium, and this results in the facile tendency of marginal exposed Ni(II) species being oxidized into the OER active Ni(III) species. Such charge transfer behaviors are further evidenced by the high-resolution Cr2+ and Fe2+ spectra [53]. As illustrated in Figure 3c, the deconvoluted Cr2+ 3p3/2 peak shows a doublet, in which the binding energy at 578.0 eV and 576.9 eV is ascribed to chromium hydroxide (Cr-OH) and chromium oxide (Cr-O), respectively. Both values are slightly higher than that from NiCr-LDH, implying that Cr (III) gains fewer electrons from Ni (II) in NiFe@NiCr-LDH. Similarly, the binding energy of the Fe 2p3/2 electrons (Figure 3d) decreased by ~0.4 eV for NiFe@NiCr-LDH (711.3 eV) as compared to that of NiFe-LDH (711.7 eV). Previous studies have revealed that such rational electron transfer from the low valent metals to the high valent metals can cause charge distribution in the catalyst and is hence favorable for facilitating the electrocatalytic kinetics [49].

3.2. Electrochemical Performance toward OER

The electrochemical performance of the catalysts toward oxygen evolution reaction (OER) was performed next in O2-saturated 1.0 M KOH aqueous solution [54]. As presented in Figure 4a, the polarization curves of the series of samples distinguish their OER activities. To afford a current density of 10 mA cm−2, the required overpotential is 266 mV for NiFe@NiCr-LDH, which is close to that of the benchmark RuO2 (249 mV) catalyst. This value is also much lower than that of NiCr-LDH (327 mV) and NiCr@NiFe-LDH (340 mV), indicating more active sites on the external NiFe-LDH phase instead of the inner NiCr-LDH phase. The overpotential value of NiFe@NiCr-LDH is also smaller than that of NiCrFe-LTH (282 mV) and NiFe-LDH (306 mV), indicating that the introduction of Cr(III) and its location are both crucial for enhancing the OER performance. The reaction kinetics were then analyzed by Tafel plots, where the Tafel slope values were extrapolated (Figure 4b). NiFe@NiCr-LDH has a small Tafel slope of 63 mV dec−1, which is lower than that of NiCrFe-LTH (68 mV dec−1) and NiFe-LDH (70 mV dec−1), indicating rapid OER kinetics. The Tafel slope values of NiCr-LDH (104 mV dec−1) and NiCr@NiFe-LDH (99 mV dec−1) are higher because of the limitation from the external NiCr-LDH phase. Furthermore, the OER activity of NiFe@NiCr-LDH outperforms or is at least comparable to recently reported LDH catalysts with a similar brucite-like layered structure under the same reaction conditions, as summarized in Table S2 [38,44,47,48,49,50,51,52,53,54,55].
To probe the underlying intrinsic reason for the brilliant OER performance of NiFe@NiCr-LDH, the electrochemically active surface area (ECSA) of all the samples was then evaluated (Figures S8 and S9) [34]. It is known that the ESCA value is linearly related to the double layer capacitance (Cdl) in the non-Faradaic potential region. As presented in Figure 4c, NiFe@NiCr-LDH has a Cdl value of 0.45 mF cm−2, which is larger than that of the other samples, indicating more exposed active sites for electrocatalysis. Moreover, the reaction kinetics was also examined by electron impedance spectroscopy (EIS), where the charge transfer behaviors during the OER process can be appraised. As illustrated in Figure 4d, NiFe@NiCr-LDH has a quite small semi-cycle in the series, suggesting rather fast reaction kinetics. By using an equivalent circuit, the EIS spectra are fitted and shown in the inset of Figure 4d, and all the calculated values are compiled in Table S3. NiFe@NiCr-LDH has the smallest charge transfer resistivity, indicating the most favorable charge transfer capability. In addition, the corresponding fitting Cdl value of NiFe@NiCr-LDH (0.44 mF cm−2) is also the highest among all the samples, and such high capacitance can facilitate the OER process as well.
Long-term durability is another critical parameter for assessing the intrinsic electrocatalytic performance of the catalyst for prolonged operation [55]. As depicted in Figure 4e, after about 20 h of continuous chronoamperometric testing, the initial current density of NiFe@NiCr-LDH only decreased by about ~11%, while, in stark contrast, NiCr@NiFe-LDH dropped by over 50%, and RuO2 also decreased by ~40%. That main location of the metal elements also affects the samples’ OER durability. The markedly outperformed long-term stability of NiFe@NiCr-LDH was further validated by the accelerated durability test (ADT). As shown in Figure 4f, after 5000 cycles of potential scanning, a negligible potential shift at the current density of 10 mA cm−2 was observed for NiFe@NiCr-LDH, while NiCr-LDH exhibited a positive overpotential shift of 24 mV. In addition, the EIS spectra before and after the OER test can further verify the outstanding stability of NiFe@NiCr-LDH [17]. As presented in Figure S10 and Table S3, the charge transfer kinetics remained almost unchanged after the OER test in NiFe@NiCr-LDH, but for the other samples, much more sluggish charge transfer kinetics were observed. Moreover, the corresponding fitting Cdl value of NiFe@NiCr-LDH exhibited a negligible change (0.43 mF cm−2), but for the other samples, the Cdl values decreased substantially. In other words, the exposed outer NiCr-LDH phases would reduce the samples’ OER long-term durability.
To further probe the physical origin of the outstanding activity and stability of the NiFe@NiCr-LDH catalyst, XPS analysis was conducted before and after the OER test [43]. As depicted in Figure 5a, the binding energy of both the Ni2+ 2p1/2 and Ni2+ 2p3/2 electrons in NiFe@NiCr-LDH shifted positively after the OER test, and more importantly, Ni3+ species also appeared, while in sharp contrast, for NiCr@NiFe-LDH (Figure 5b), the binding energy of both the Ni2+ 2p1/2 and Ni2+ 2p3/2 electrons exhibited a negligible change without the presence of Ni3+ species. This indicates that, in the presence of Fe3+, the Ni2+ species becomes partially oxidized in the outer shell in NiFe@NiCr-LDH; however, the Cr3+ species is not able to promote the oxidation of Ni2+ in the outer shell of NiCr@NiFe-LDH. Meanwhile, despite the binding energy of the Cr(-O) and Cr(-OH) 2p electrons remaining almost unchanged after the OER process in both NiFe@NiCr-LDH (Figure 5c) and NiCr@NiFe-LDH (Figure 5d), the Cr6+ species is somehow present in NiCr@NiFe-LDH but cannot be found in NiFe@NiCr-LDH [51]. In other words, Cr3+, instead of Ni2+ species, becomes partially oxidized in the outer shell in NiCr@NiFe-LDH. In addition, as described in Figure S11, the binding energy of both the Fe3+ 2p3/2 and Fe 2p1/2 electrons increased over 1 eV in NiFe@NiCr-LDH, while the shift in NiCr@NiFe-LDH is negligible, suggesting that Fe3+ species in different locations (inner or outer shell) play markedly different roles [50]. When the NiCr phase is exposed (NiCr@NiFe-LDH), Cr(III) becomes easily oxidized into Cr(VI), but Ni still maintains +2 valence; hence, the OER activity cannot be promoted, while, in contrast, the Fe(III) species in the outer shell (NiFe@NiCr-LDH) facilitate the oxidation of Ni(II) into active Ni(III), and Cr(III) is still well-preserved during the OER process to maintain the good conductivity. Such synergistic effect (outer NiFe-LDH phases provide activity; inner NiCr-LDH phases guarantee stability) boosts the NiFe@NiCr-LDH’s OER performance.

4. Conclusions

In summary, we developed an NiFe@NiCr-LDH catalyst with heterogeneous interfaces for alkaline OER. Multiple characterizations revealed that a well-defined heterostructure is formed at the interface between the outer NiFe-LDH phase and the inner NiCr-LDH phase. In 1 M KOH, NiFe@NiCr-LDH showed excellent OER performance, evidenced by a low overpotential at 10 mA cm−2 and a small Tafel slope, both close to those of the benchmark RuO2 catalyst, as well as markedly superior long-term stability compared to that of RuO2. Such impressive performance is attributed to the fact that the Fe(III) species in the outer shell accelerate the formation of Ni(III) species as active sites for OER in NiFe@NiCr-LDH, which cannot be achieved for NiCr@NiFe-LDH and other samples without a heterostructure. This study provides a facile strategy for constructing an NiFe-LDH-based heterostructure for OER and other electrochemical energy technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16082968/s1.

Author Contributions

S.L., Y.T., C.G. and Y.L. conducted the literature search and wrote the draft. S.L. and Z.T. conceived the idea. S.L. and Z.T. polished the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Research Fund Program of the State Key Laboratory of Subtropical Building Science at South China University of Technology (Grant No. 2022ZB02). Z. T. acknowledges the financial support from Guangdong Natural Science Funds (No. 2022A1515011840). Y. L. gratefully acknowledges the financial support from the Basic Research and Applied Basic Research Foundation of Guangzhou (202201010650).

Data Availability Statement

The authors will supply the relevant data in response to reasonable requests.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, T.; Hu, T.; Dai, L.; Li, C.M. Metal-free photo- and electro-catalysts for hydrogen evolution reaction. J. Mater. Chem. A 2020, 8, 23674–23698. [Google Scholar] [CrossRef]
  2. Ma, Y.; Ma, Y.; Wang, Q.; Schweidler, S.; Botros, M.; Fu, T.; Hahn, H.; Brezesinski, T.; Breitung, B. High-entropy energy materials: Challenges and new opportunities. Energy Environ. Sci. 2021, 14, 2883–2905. [Google Scholar] [CrossRef]
  3. Yang, Y.; Yu, Y.; Li, J.; Chen, Q.; Du, Y.; Rao, P.; Li, R.; Jia, C.; Kang, Z.; Deng, P.; et al. Engineering Ruthenium-Based Electrocatalysts for Effective Hydrogen Evolution Reaction. Nano-Micro Lett. 2021, 13, 160. [Google Scholar] [CrossRef] [PubMed]
  4. Qi, J.; Zhang, W.; Cao, R. Solar-to-Hydrogen Energy Conversion Based on Water Splitting. Adv. Energy Mater. 2018, 8, 1701620. [Google Scholar] [CrossRef]
  5. Zhou, B.; Gao, R.; Zou, J.J.; Yang, H. Surface Design Strategy of Catalysts for Water Electrolysis. Small 2022, 18, 2202336. [Google Scholar] [CrossRef]
  6. Zhang, J.; Zhang, Q.; Feng, X. Support and Interface Effects in Water-Splitting Electrocatalysts. Adv. Mater. 2019, 31, 1808167. [Google Scholar] [CrossRef]
  7. Li, H.; Pan, Y.; Lai, J.; Wang, L.; Feng, S. Strong High Entropy Alloy-Support Interaction Enables Efficient Electrocatalytic Water Splitting at High Current Density. Chin. J. Struct. Chem. 2022, 41, 2208003–2208011. [Google Scholar]
  8. Wang, F.-G.; Liu, X.; Lv, Q.-X.; Liu, B.; Chai, Y.-M.; Dong, B. Transition Metal Boride-Based Materials for Electrocatalytic Water Splitting. Chin. J. Struct. Chem. 2022, 41, 2209008–2209044. [Google Scholar]
  9. Suen, N.T.; Hung, S.F.; Quan, Q.; Zhang, N.; Xu, Y.J.; Chen, H.M. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef]
  10. Feng, C.; Faheem, M.B.; Fu, J.; Xiao, Y.; Li, C.; Li, Y. Fe-Based Electrocatalysts for Oxygen Evolution Reaction: Progress and Perspectives. ACS Catal. 2020, 10, 4019–4047. [Google Scholar] [CrossRef]
  11. Rao, R.R.; Kolb, M.J.; Halck, N.B.; Pedersen, A.F.; Mehta, A.; You, H.; Stoerzinger, K.A.; Feng, Z.; Hansen, H.A.; Zhou, H.; et al. Towards identifying the active sites on RuO2(110) in catalyzing oxygen evolution. Energy Environ. Sci. 2017, 10, 2626–2637. [Google Scholar] [CrossRef] [Green Version]
  12. Binninger, T.; Doublet, M.L. The Ir-OOOO-Ir transition state and the mechanism of the oxygen evolution reaction on IrO2(110). Energy Environ. Sci. 2022, 15, 2519–2528. [Google Scholar] [CrossRef]
  13. Liu, Y.; Liang, X.; Chen, H.; Gao, R.; Shi, L.; Yang, L.; Zou, X. Iridium-containing water-oxidation catalysts in acidic electrolyte. Chin. J. Catal. 2021, 42, 1054–1077. [Google Scholar] [CrossRef]
  14. Shan, J.; Guo, C.; Zhu, Y.; Chen, S.; Song, L.; Jaroniec, M.; Zheng, Y.; Qiao, S.-Z. Charge-Redistribution-Enhanced Nanocrystalline Ru@IrOx Electrocatalysts for Oxygen Evolution in Acidic Media. Chem 2019, 5, 445–459. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, C.; Jin, L.; Shang, H.; Xu, H.; Shiraishi, Y.; Du, Y. Advances in engineering RuO2 electrocatalysts towards oxygen evolution reaction. Chin. Chem. Lett. 2021, 32, 2108–2116. [Google Scholar] [CrossRef]
  16. Yu, J.; He, Q.; Yang, G.; Zhou, W.; Shao, Z.; Ni, M. Recent Advances and Prospective in Ruthenium-Based Materials for Electrochemical Water Splitting. ACS Catal. 2019, 9, 9973–10011. [Google Scholar] [CrossRef]
  17. Sun, H.; Tung, C.W.; Qiu, Y.; Zhang, W.; Wang, Q.; Li, Z.; Tang, J.; Chen, H.C.; Wang, C.; Chen, H.M. Atomic Metal-Support Interaction Enables Reconstruction-Free Dual-Site Electrocatalyst. J. Am. Chem. Soc. 2022, 144, 1174–1186. [Google Scholar] [CrossRef] [PubMed]
  18. Zhou, Y.; Li, J.; Gao, X.; Chu, W.; Gao, G.; Wang, L.-W. Recent advances in single-atom electrocatalysts supported on two-dimensional materials for the oxygen evolution reaction. J. Mater. Chem. A 2021, 9, 9979–9999. [Google Scholar] [CrossRef]
  19. Wei, M.; Li, J.; Chu, W.; Wang, N. Phase control of 2D binary hydroxides nanosheets via controlling-release strategy for enhanced oxygen evolution reaction and supercapacitor performances. J. Energy Chem. 2019, 38, 26–33. [Google Scholar] [CrossRef] [Green Version]
  20. Fu, C.-L.; Wang, Y.; Huang, J.-H. Hybrid of Quaternary Layered Double Hydroxides and Carbon Nanotubes for Oxygen Evolution Reaction. Chin. J. Struct. Chem. 2020, 39, 1807–1816. [Google Scholar]
  21. Luo, M.; Cai, Z.; Wang, C.; Bi, Y.; Qian, L.; Hao, Y.; Li, L.; Kuang, Y.; Li, Y.; Lei, X.; et al. Phosphorus oxoanion-intercalated layered double hydroxides for high-performance oxygen evolution. Nano Res. 2017, 10, 1732–1739. [Google Scholar] [CrossRef]
  22. 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]
  23. Rajeshkhanna, G.; Singh, T.I.; Kim, N.H.; Lee, J.H. Remarkable Bifunctional Oxygen and Hydrogen Evolution Electrocatalytic Activities with Trace-Level Fe Doping in Ni- and Co-Layered Double Hydroxides for Overall Water-Splitting. ACS Appl. Mater. Interfaces 2018, 10, 42453–42468. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Qiao, M.; Li, Y.; Wang, S. Tuning Surface Electronic Configuration of NiFe LDHs Nanosheets by Introducing Cation Vacancies (Fe or Ni) as Highly Efficient Electrocatalysts for Oxygen Evolution Reaction. Small 2018, 14, 1800136. [Google Scholar] [CrossRef] [PubMed]
  25. Zhou, D.; Cai, Z.; Bi, Y.; Tian, W.; Luo, M.; Zhang, Q.; Zhang, Q.; Xie, Q.; Wang, J.; Li, Y.; et al. Effects of redox-active interlayer anions on the oxygen evolution reactivity of NiFe-layered double hydroxide nanosheets. Nano Res. 2018, 11, 1358–1368. [Google Scholar] [CrossRef]
  26. Mohammed-Ibrahim, J. A review on NiFe-based electrocatalysts for efficient alkaline oxygen evolution reaction. J. Power Sources 2020, 448, 227375. [Google Scholar] [CrossRef]
  27. Li, X.; Zai, J.; Liu, Y.; He, X.; Xiang, S.; Ma, Z.; Qian, X. Atomically thin layered NiFe double hydroxides assembled 3D microspheres with promoted electrochemical performances. J. Power Sources 2016, 325, 675–681. [Google Scholar] [CrossRef]
  28. Ma, W.; Ma, R.; Wang, C.; Liang, J.; Liu, X.; Zhou, K.; Sasaki, T. A Superlattice of Alternately Stacked Ni-Fe Hydroxide Nanosheets and Graphene for Efficient Splitting of Water. ACS Nano 2015, 9, 1977–1984. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, A.-L.; Xu, H.; Li, G.-R. NiCoFe Layered Triple Hydroxides with Porous Structures as High-Performance Electrocatalysts for Overall Water Splitting. ACS Energy Lett. 2016, 1, 445–453. [Google Scholar] [CrossRef]
  30. Liu, S.; Wan, R.; Lin, Z.; Liu, Z.; Liu, Y.; Tian, Y.; Qin, D.-D.; Tang, Z. Probing the Co role in promoting the OER and Zn–air battery performance of NiFe-LDH: A combined experimental and theoretical study. J. Mater. Chem. A 2022, 10, 5244–5254. [Google Scholar] [CrossRef]
  31. Chen, G.; Wang, T.; Zhang, J.; Liu, P.; Sun, H.; Zhuang, X.; Chen, M.; Feng, X. Accelerated Hydrogen Evolution Kinetics on NiFe-Layered Double Hydroxide Electrocatalysts by Tailoring Water Dissociation Active Sites. Adv. Mater. 2018, 30, 1706279. [Google Scholar] [CrossRef] [PubMed]
  32. Jin, Y.; Huang, S.; Yue, X.; Du, H.; Shen, P.K. Mo- and Fe-Modified Ni(OH)2/NiOOH Nanosheets as Highly Active and Stable Electrocatalysts for Oxygen Evolution Reaction. ACS Catal. 2018, 8, 2359–2363. [Google Scholar] [CrossRef]
  33. Zhang, J.; Liu, J.; Xi, L.; Yu, Y.; Chen, N.; Sun, S.; Wang, W.; Lange, K.M.; Zhang, B. Single-Atom Au/NiFe Layered Double Hydroxide Electrocatalyst: Probing the Origin of Activity for Oxygen Evolution Reaction. J. Am. Chem. Soc. 2018, 140, 3876–3879. [Google Scholar] [CrossRef]
  34. Li, P.; Wang, M.; Duan, X.; Zheng, L.; Cheng, X.; Zhang, Y.; Kuang, Y.; Li, Y.; Ma, Q.; Feng, Z.; et al. Boosting oxygen evolution of single-atomic ruthenium through electronic coupling with cobalt-iron layered double hydroxides. Nat. Commun. 2019, 10, 1711. [Google Scholar] [CrossRef] [Green Version]
  35. Xu, S.-J.; Zhou, Y.-N.; Shen, G.-P.; Dong, B. Ni(OH)2 Derived from NiS2 Induced by Reflux Playing Three Roles for Hydrogen/Oxygen Evolution Reaction. Chin. J. Struct. Chem. 2022, 41, 2208052–2208057. [Google Scholar]
  36. Wu, Y.; Song, M.; Huang, Y.-C.; Dong, C.-L.; Li, Y.; Lu, Y.; Zhou, B.; Wang, D.; Jia, J.; Wang, S.; et al. Promoting surface reconstruction of NiFe layered double hydroxides via intercalating [Cr(C2O4)3]3− for enhanced oxygen evolution. J. Energy Chem. 2022, 74, 140–148. [Google Scholar] [CrossRef]
  37. Yao, Y.; Xu, Z.; Cheng, F.; Li, W.; Cui, P.; Xu, G.; Xu, S.; Wang, P.; Sheng, G.; Yan, Y.; et al. Unlocking the potential of graphene for water oxidation using an orbital hybridization strategy. Energy Environ. Sci. 2018, 11, 407–416. [Google Scholar] [CrossRef]
  38. Babu, S.P.; Falch, A. Recent Developments on Cr-Based Electrocatalysts for the Oxygen Evolution Reaction in Alkaline Media. ChemCatChem 2022, 14, e202200364. [Google Scholar] [CrossRef]
  39. Liu, T.; Diao, P. Nickel foam supported Cr-doped NiCo2O4/FeOOH nanoneedle arrays as a high-performance bifunctional electrocatalyst for overall water splitting. Nano Res. 2020, 13, 3299–3309. [Google Scholar] [CrossRef]
  40. Sun, Z.; Yuan, M.; Yang, H.; Lin, L.; Jiang, H.; Ge, S.; Li, H.; Sun, G.; Ma, S.; Yang, X. 3D Porous Amorphous gamma-CrOOH on Ni Foam as Bifunctional Electrocatalyst for Overall Water Splitting. Inorg. Chem. 2019, 58, 4014–4018. [Google Scholar] [CrossRef]
  41. Dong, C.; Zhang, X.; Xu, J.; Si, R.; Sheng, J.; Luo, J.; Zhang, S.; Dong, W.; Li, G.; Wang, W.; et al. Ruthenium-Doped Cobalt-Chromium Layered Double Hydroxides for Enhancing Oxygen Evolution through Regulating Charge Transfer. Small 2020, 16, 1905328. [Google Scholar] [CrossRef] [PubMed]
  42. Dong, C.; Yuan, X.; Wang, X.; Liu, X.; Dong, W.; Wang, R.; Duan, Y.; Huang, F. Rational design of cobalt–chromium layered double hydroxide as a highly efficient electrocatalyst for water oxidation. J. Mater. Chem. A 2016, 4, 11292–11298. [Google Scholar] [CrossRef]
  43. 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 Environ. 2020, 272, 118959. [Google Scholar] [CrossRef]
  44. Yang, Y.; Dang, L.; Shearer, M.J.; Sheng, H.; Li, W.; Chen, J.; Xiao, P.; Zhang, Y.; Hamers, R.J.; Jin, S. Highly Active Trimetallic NiFeCr Layered Double Hydroxide Electrocatalysts for Oxygen Evolution Reaction. Adv. Energy Mater. 2018, 8, 1703189. [Google Scholar] [CrossRef]
  45. Ye, W.; Fang, X.; Chen, X.; Yan, D. A three-dimensional nickel-chromium layered double hydroxide micro/nanosheet array as an efficient and stable bifunctional electrocatalyst for overall water splitting. Nanoscale 2018, 10, 19484–19491. [Google Scholar] [CrossRef]
  46. Xiang, Q.; Li, F.; Chen, W.; Ma, Y.; Wu, Y.; Gu, X.; Qin, Y.; Tao, P.; Song, C.; Shang, W.; et al. In Situ Vertical Growth of Fe–Ni Layered Double-Hydroxide Arrays on Fe–Ni Alloy Foil: Interfacial Layer Enhanced Electrocatalyst with Small Overpotential for Oxygen Evolution Reaction. ACS Energy Lett. 2018, 3, 2357–2365. [Google Scholar] [CrossRef]
  47. Chen, S.; Yu, C.; Cao, Z.; Huang, X.; Wang, S.; Zhong, H. Trimetallic NiFeCr-LDH/MoS2 composites as novel electrocatalyst for OER. Int. J. Hydrog. Energy 2021, 46, 7037–7046. [Google Scholar] [CrossRef]
  48. Bo, X.; Li, Y.; Hocking, R.K.; Zhao, C. NiFeCr Hydroxide Holey Nanosheet as Advanced Electrocatalyst for Water Oxidation. ACS Appl. Mater. Interfaces 2017, 9, 41239–41245. [Google Scholar] [CrossRef]
  49. Zhong, H.; Liu, T.; Zhang, S.; Li, D.; Tang, P.; Alonso-Vante, N.; Feng, Y. Template-free synthesis of three-dimensional NiFe-LDH hollow microsphere with enhanced OER performance in alkaline media. J. Energy Chem. 2019, 33, 130–137. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, Y.; Cui, X.; Gao, D.; He, H.; Ou, Y.; Zhou, M.; Lai, Q.; Wei, X.; Xiao, P.; Zhang, Y. Trimetallic CoFeCr hydroxide electrocatalysts synthesized at a low temperature for accelerating water oxidationviatuning the electronic structure of active sites. Sustain. Energ. Fuels 2020, 4, 3647–3653. [Google Scholar] [CrossRef]
  51. Bo, X.; Li, Y.; Chen, X.; Zhao, C. High valence chromium regulated cobalt-iron-hydroxide for enhanced water oxidation. J. Power Sources 2018, 402, 381–387. [Google Scholar] [CrossRef]
  52. Thenuwara, A.C.; Attanayake, N.H.; Yu, J.; Perdew, J.P.; Elzinga, E.J.; Yan, Q.; Strongin, D.R. Cobalt Intercalated Layered NiFe Double Hydroxides for the Oxygen Evolution Reaction. J. Phys. Chem. B 2018, 122, 847–854. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, M.H.; Lou, Z.X.; Wu, X.; Liu, Y.; Zhao, J.Y.; Sun, K.Z.; Li, W.X.; Chen, J.; Yuan, H.Y.; Zhu, M.; et al. Operando High-Valence Cr-Modified NiFe Hydroxides for Water Oxidation. Small 2022, 18, 2200303. [Google Scholar] [CrossRef] [PubMed]
  54. Yu, L.; Yang, J.F.; Guan, B.Y.; Lu, Y.; Lou, X.W.D. Hierarchical Hollow Nanoprisms Based on Ultrathin Ni-Fe Layered Double Hydroxide Nanosheets with Enhanced Electrocatalytic Activity towards Oxygen Evolution. Angew. Chem. Int. Ed. 2018, 130, 178–182. [Google Scholar] [CrossRef]
  55. Yan, L.; Du, Z.; Lai, X.; Lan, J.; Liu, X.; Liao, J.; Feng, Y.; Li, H. Synergistically modulating the electronic structure of Cr-doped FeNi LDH nanoarrays by O-vacancy and coupling of MXene for enhanced oxygen evolution reaction. Int. J. Hydrog. Energy 2023, 48, 1892–1903. [Google Scholar] [CrossRef]
Scheme 1. (a) Schematic illustration for synthesizing NiFe@NiCr-LDH, where cyan, indigo-blue, yellow, purple, red, green, white, and gray atoms represent Ni, Cr, Fe, F, O, N, C, and H, respectively; Ideal models for (b) NiCr-LDH and (c) NiFe@NiCr-LDH with interlayer anions and H2O molecules.
Scheme 1. (a) Schematic illustration for synthesizing NiFe@NiCr-LDH, where cyan, indigo-blue, yellow, purple, red, green, white, and gray atoms represent Ni, Cr, Fe, F, O, N, C, and H, respectively; Ideal models for (b) NiCr-LDH and (c) NiFe@NiCr-LDH with interlayer anions and H2O molecules.
Materials 16 02968 sch001
Figure 1. (a) XRD patterns of NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, and NiFe-LDH, with FWHM values of LDHs’ (003) peaks; (b) FTIR spectra of the as-prepared LDH samples; (c) Typical SEM images of NiFe-LDH/NiCr-LDH; (df) Typical SEM images of NiFe@NiCr-LDH with different magnifications.
Figure 1. (a) XRD patterns of NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, and NiFe-LDH, with FWHM values of LDHs’ (003) peaks; (b) FTIR spectra of the as-prepared LDH samples; (c) Typical SEM images of NiFe-LDH/NiCr-LDH; (df) Typical SEM images of NiFe@NiCr-LDH with different magnifications.
Materials 16 02968 g001
Figure 2. (a) Typical TEM image, (b) HR-TEM images along with lattice fringes of NiCr-LDH and NiFe-LDH phases in NiFe@NiCr-LDH, (c) SAED pattern of NiFe@NiCr-LDH nanosheets, (d) Typical HR-TEM images of NiCr@NiFe-LDH nanosheets along with lattice fringes of NiCr-LDH and NiFe-LDH phases, (e) HAADF-STEM image of an NiFe@NiCr-LDH nanosphere, (fi) Elemental mapping of Ni, Cr, Fe, and O elements of the NiFe@NiCr-LDH nanosphere.
Figure 2. (a) Typical TEM image, (b) HR-TEM images along with lattice fringes of NiCr-LDH and NiFe-LDH phases in NiFe@NiCr-LDH, (c) SAED pattern of NiFe@NiCr-LDH nanosheets, (d) Typical HR-TEM images of NiCr@NiFe-LDH nanosheets along with lattice fringes of NiCr-LDH and NiFe-LDH phases, (e) HAADF-STEM image of an NiFe@NiCr-LDH nanosphere, (fi) Elemental mapping of Ni, Cr, Fe, and O elements of the NiFe@NiCr-LDH nanosphere.
Materials 16 02968 g002
Figure 3. High-resolution XPS spectra of the (a) O 1s, (b) Ni 2p, (c) Cr 2p, and (d) Fe 2p electrons in NiFe@NiCr-LDH, NiCr-LDH, and NiFe-LDH.
Figure 3. High-resolution XPS spectra of the (a) O 1s, (b) Ni 2p, (c) Cr 2p, and (d) Fe 2p electrons in NiFe@NiCr-LDH, NiCr-LDH, and NiFe-LDH.
Materials 16 02968 g003
Figure 4. (a) LSV polarization curves and the corresponding Tafel plots (b) of NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, NiFe-LDH, and RuO2 in 1.0 M KOH; (c) The current density at 1.18 V vs. the scan rate for NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, and NiFe-LDH; (d) Electrochemical impedance spectra (EIS) of the samples; (e) The i-t curve at the current density of 10 mA cm−2 for 24 h; (f) The LSV polarization curves of the NiFe@NiCr-LDH and NiCr-LDH catalyst before and after 5000 CV cycles.
Figure 4. (a) LSV polarization curves and the corresponding Tafel plots (b) of NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, NiFe-LDH, and RuO2 in 1.0 M KOH; (c) The current density at 1.18 V vs. the scan rate for NiFe@NiCr-LDH, NiCr@NiFe-LDH, NiCrFe-LTH, NiCr-LDH, and NiFe-LDH; (d) Electrochemical impedance spectra (EIS) of the samples; (e) The i-t curve at the current density of 10 mA cm−2 for 24 h; (f) The LSV polarization curves of the NiFe@NiCr-LDH and NiCr-LDH catalyst before and after 5000 CV cycles.
Materials 16 02968 g004
Figure 5. High-resolution XPS spectra of the Ni 2p electrons in (a) NiFe@NiCr-LDH and (b) NiCr@NiFe-LDH before and after the OER test; High-resolution XPS spectra of the Cr 2p electrons in (c) NiFe@NiCr-LDH and (d) NiCr@NiFe-LDH before and after the OER test.
Figure 5. High-resolution XPS spectra of the Ni 2p electrons in (a) NiFe@NiCr-LDH and (b) NiCr@NiFe-LDH before and after the OER test; High-resolution XPS spectra of the Cr 2p electrons in (c) NiFe@NiCr-LDH and (d) NiCr@NiFe-LDH before and after the OER test.
Materials 16 02968 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, S.; Tang, Y.; Guo, C.; Liu, Y.; Tang, Z. Heterostructure of NiFe@NiCr-LDH for Active and Durable Oxygen Evolution Reactions in Alkaline Media. Materials 2023, 16, 2968. https://doi.org/10.3390/ma16082968

AMA Style

Liu S, Tang Y, Guo C, Liu Y, Tang Z. Heterostructure of NiFe@NiCr-LDH for Active and Durable Oxygen Evolution Reactions in Alkaline Media. Materials. 2023; 16(8):2968. https://doi.org/10.3390/ma16082968

Chicago/Turabian Style

Liu, Sanchuan, Yujun Tang, Chengyu Guo, Yonggang Liu, and Zhenghua Tang. 2023. "Heterostructure of NiFe@NiCr-LDH for Active and Durable Oxygen Evolution Reactions in Alkaline Media" Materials 16, no. 8: 2968. https://doi.org/10.3390/ma16082968

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