Download PDF
Article  |  Open Access  |  12 Dec 2021

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillar electrode

Views: 26819 |  Downloads: 3657 |  Cited:   17
Energy Mater 2021;1:100015.
10.20517/energymater.2021.16 |  © The Author(s) 2021.
Author Information
Article Notes
Cite This Article

Abstract

Seawater splitting powered by solar or wind sources is a significant renewable energy storage technology for the production of green hydrogen energy. However, both the chlorine evolution reaction and chloride corrosion are intractable issues in seawater splitting. Here, a porous electrode based on a phosphate-intercalated NiFe (oxy)hydroxide shell coated on a nickel molybdate (NiMoO4) micropillar core (denoted as P-NiFe@NiMoO4) is synthesized through an electrochemical oxidation strategy. During the electrochemical oxidation process, the etching of MoO2 promotes the reconstruction of NiFe (oxy)hydroxide and the formation of porous structures in an alkaline solution. The optimized P-NiFe@NiMoO4 electrocatalysts afford a low overpotential of 258 mV at a current density of 100 mA/cm2 in alkaline seawater. By pairing the anode with a cathode of as-synthesized P-NiMoO, the electrolyzer presents a low voltage of 1.63 V at 100 mA/cm2 in alkaline seawater with excellent stability. Moreover, the remarkable stability of the anode seems to be attributed to the in-situ phosphate formed during the electrochemical oxidation process to passivate chloride corrosion.

Keywords

NiFe (oxy)hydroxide, NiMoO4 micropillars, oxygen evolution reaction, seawater splitting

INTRODUCTION

With the depletion of energy resources and environmental destruction, the exploration of new energy sources has become an urgent priority[1-3]. Hydrogen is the optimal clean energy as a result of its high gravimetric energy density (142 MJ/kg) and pollution-free application[4-8]. Although water splitting is an efficient approach to yield hydrogen, the oxygen evolution reaction (OER) represents a bottleneck owing to its sluggish four-electron process[9-14]. Recently, considering the shortage of freshwater, there has been a dramatic increase in interest in seawater splitting because seawater is the most abundant resource globally. Nevertheless, several challenges remain before this technology can become competitive.

Excess chloride anions represent a crucial challenge during seawater splitting and they may occur in the competitive chlorine evolution reaction at the anode[15,16]. In an alkaline solution, the chlorine furthers react with OH- to form ClO- with an onset potential that is ~490 mV higher than that of the OER. Hence, the promising overpotential of OER electrocatalysts under high current must be kept well below 490 mV, under which a seawater electrolysis efficiency of 100% can be achieved. Furthermore, the corrosion of catalysts and substrates from aggressive chloride anions in seawater is another challenge and can lead to poor long-term stability[17]. To slow down this corrosion process, it is vital to develop corrosion-resistant electrodes for seawater splitting.

Traditionally, Pt-, Ir- and Ru-based noble metal catalysts are used for water splitting because of their relatively good activity for both the hydrogen evolution reaction (HER) and OER[18-21]. Nevertheless, considering the high cost, scarcity and poor stability of noble metal catalysts, earth-abundant catalysts, such as transition-metal oxides[22,23], (oxy)hydroxides[24-26], phosphides[27-29], nitrides[15,30,31], borides[32], sulfides[33,34] and selenides[35,36], have been used as promising alternatives in recent years. In particular, transition-metal (oxy)hydroxides have attracted considerable attention for seawater electrolysis due to their high activity. For example, a NiFe layered double hydroxide catalyst synthesized by Dionigi et al.[37] showed a high overpotential of 359 mV at 10 mA/cm2 in a simulated seawater electrolyte. More recently, Kuang et al.[17] reported that an anode based on a NiFe hydroxide-coated NiSx layer could achieve a high current density of 400 mA/cm2 at an overpotential of 510 mV, which tremendously boosted the development of hydrogen production. Liu et al.[38] recently constructed a Fe-doped NiOOH catalyst by self-reconstruction, which exhibited an overpotential of 292 mV at 100 mA/cm2 and excellent stability. Furthermore, the construction of highly efficient seawater splitting electrocatalysts is essential for balancing catalytic activity and stability.

Herein, we report a porous electrode based on a phosphate-intercalated NiFe (oxy)hydroxide shell coated on a nickel molybdate (NiMoO4) micropillar core (denoted as P-NiFe@NiMoO4) on nickel foam (NF) by an electrochemical oxidation strategy for alkaline seawater splitting. The P-NiFe@NiMoO4 core-shell electrode requires low overpotentials of 238 and 258 mV at a current density of 100 mA/cm2 in 1 M KOH and 1 M KOH + seawater electrolytes, respectively. Moreover, the overall alkaline seawater splitting electrolyzer, in which P-NiFe@NiMoO4 is coupled with another efficient HER electrode of P-NiMoO, requires a low voltage of 1.63 V at a current density of 100 mA/cm2, benefiting from the distinct porosity and large surface area. Along with excellent catalytic performance, the two-electrode electrolyzer also shows superior stability with a slight decrease over 200 h in alkaline seawater at ambient temperature, attributed to the formation of phosphate that intercalates into NiFe (oxy)hydroxide during the reconstruction process to prevent Cl- corrosion.

EXPERIMENTAL

Reagents

FeSO4·9H2O [analytical reagent (AR)], (NH4)6Mo7O24·4H2O (AR), Ni(NO3)2·6H2O (AR), NaH2PO2 (AR), KOH (AR), HCl (AR), ethanol (AR) and commercial RuO2 were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd, China. With a thickness of 1.6 mm and a porosity of ~95%, NF was provided by Suzhou Jiashide Metal Foam Co., Ltd. Deionized (DI) water was obtained from a Millipore Milli-Q purification system.

Preparation of catalysts

Synthesis of FeNiMoO and NiMoO precursors on nickel foam

A piece of NF (2 × 4 cm2) was cleaned by ultrasonication with 2 M HCl for 15 min and the substrate was then washed with ethanol and DI water several times. A mixture of 378.3 mg of Ni(NO3)2·6H2O, 202 mg of FeSO4·7H2O and 395.5 mg of (NH4)6Mo7O24·4H2O was dissolved in 40 mL of DI water to form a homogeneous solution. The solution and as-prepared NF substrate were then transferred to a 100 mL Teflon-lined stainless-steel autoclave, which was heated at 150 °C for 6 h. After cooling to room temperature naturally, the sample was washed with DI water several times. The NiMoO precursor was prepared similarly without the addition of FeSO4·9H2O.

Synthesis of P-FeNiMoO, P-NiMoO and A-FeNiMoO electrodes

A total of 400 mg of NaH2PO2 on porcelain was put upstream of a furnace and the as-prepared FeNiMoO precursor was put downstream. The furnace was heated at 450 °C for 2 h under an N2 atmosphere. Similarly, the P-NiMoO electrode was synthesized by replacing the FeNiMoO precursor with a NiMoO precursor. The A-FeNiMoO electrode was synthesized by directly annealing the FeNiMoO precursor in a tube furnace. The annealing treatment was conducted at 450 °C for 2 h under an N2 atmosphere.

Synthesis of P-NiFe@NiMoO4, A-NiFe@NiMoO4 and P-Ni@NiMoO4 electrodes

An electrochemical oxidation process was deployed to synthesize the three electrodes. The P-FeNiMoO, P-NiMoO and A-FeNiMoO electrodes were electroxidized under a current density of 100 mA/cm2 for 6 h to obtain P-NiFe@NiMoO4, P-Ni@NiMoO4 and A-NiFe@NiMoO4, respectively.

Preparation of RuO2 and Pt/C catalysts on carbon paper

To prepare the RuO2 electrode for comparison, 3 mg of RuO2 and 10 μL of Nafion were dispersed in 990 μL of ethanol and the mixture was ultrasonicated for 30 min. A 100 μL dispersion was then coated onto a clean carbon paper substrate, where the mass loading was 300 μg/cm2, followed by drying in air. Pt/C electrodes were obtained by a similar method but with a mass loading of 30 mg/cm2.

Characterization of catalysts

Physical characterization

Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy were performed with a Hitachi S-4800 scanning electron microscope equipped with an energy dispersive X-ray analyzer (HORIBA, 7593-H). Both transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) were performed with a FEI Tecnai T20 microscope. X-ray diffraction (XRD) patterns were collected on a Bruker D8 Advanced diffractometer using a Cu-Kα radiation source in the 2θ range of 10-80°. X-ray photoelectron spectroscopy (XPS) was performed on a K-Alpha X-ray photoelectron spectrometer (Thermo Scientific). Raman spectroscopy was recorded at ambient temperature on a DXR Raman microscope with a 532 nm excitation length.

Electrochemical measurements

The electrochemical performance was tested using an electrochemical workstation (CH Instruments 660E). The two half reactions of the OER and HER were each carried out at ambient temperature (25 °C) in a standard three-electrode system, with the as-prepared electrode, a graphite rod (for the HER) or Pt mesh (for the OER) and a standard Hg/HgO electrode used as the working, counter and reference electrodes, respectively. All the electrode sizes were ~1 cm-2 and the potentials were calibrated to the reversible hydrogen electrode with 85% iR compensation if not specified.

OER polarization curve measurements were performed by cyclic voltammetry (CV) at a scan rate of 5 mV/s and stability tests were carried out under 100 mA cm-2 in 1 M KOH. The HER polarization curve was collected by linear sweep voltammetry at a scan rate of 5 mV/s and stability tests were carried out under 100 mA cm-2 in 1 M KOH. Electrochemical impedance spectroscopy was measured at an overpotential of 270 mV from 0.1 Hz to 100 kHz with an amplitude of 10 mV. Cdl was measured by CV in the non-faradaic region.

The two-electrode electrolyzer was constructed by pairing the as-prepared P-NiFe@NiMoO4 with the P-NiMoO electrode as the anode and cathode, respectively. The workstation was used to measure the polarization curves in different electrolytes. The stability test was carried out with a LANHE battery tester working at a constant current density of 100 mA cm-2 at 25 °C.

RESULTS AND DISCUSSION

Scheme 1 presents a schematic for the synthesis process of the porous P-NiFe@NiMoO4 core-shell electrode, with NF used as the substrate benefiting from its natural conductivity, high surface area and abundant reserves. Firstly, a hydrothermal method was deployed to synthesize the FeNiMoO precursor. Secondly, the FeNiMoO precursor was put into a tube furnace for phosphorization treatment, followed by an electrochemical oxidation process. As shown in Supplementary Figure 1, the as-synthesized FeNiMoO precursor electrode turned from brown to black after the phosphorization process. It should be noted that the precursor turned from brown to dark yellow after an annealing treatment under an N2 atmosphere rather than phosphorization. The FeNiMoO precursor was characterized by SEM. In Supplementary Figure 2, it can be seen that the micropillars with smooth surfaces grew well on the NF. After the phosphorization process, shown in Supplementary Figure 3, the smooth surface of the P-FeNiMoO electrode was covered by a shell of dense material. In a control experiment, another electrode was synthesized with a similar procedure but heated at high temperatures without NaH2PO2 (denoted as A-FeNiMoO). In contrast, the A-FeNiMoO electrode in Supplementary Figure 4 shows that these micropillars were broken with cracks on the surface, possibly due to the absence of the covering dense material.

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO<sub>4</sub> core-shell micropillar electrode

Scheme 1. Schematic of P-NiFe@NiMoO4 electrode synthesis.

To further investigate the dense shell, a SEM image of a cross section of P-FeNiMoO is provided in Figure 1A, which clearly shows that the micropillar presents a core-shell structure. The XRD pattern of P-FeNiMoO is shown in Supplementary Figure 5. It is evident that the diffraction peaks at 26.1°, 36.7° and 53.3° could be indexed to the (110), (-211) and (220) planes of MoO2 according to the Joint Committee on Powder Diffraction Standards (JCPDS) No. 78-1072 and the diffraction peaks at 14.2°, 25.3°, 28.7°, 32.5° and 47.5° could be assigned to NiMoO4 according to JCPDS No. 86-0361, which indicates the presence of both NiMoO4 and MoO2. The XRD pattern of A-FeNiMoO was also collected and is shown in Supplementary Figure 5, where the strong peaks could be assigned to NiMoO4. Moreover, a smaller amount of MoO2 was obtained during the annealing treatment under N2 compared with the phosphorization process. Therefore, it is envisaged that the MoO2 is mainly produced from the reaction of PH3 produced from the decomposition of NaH2PO2 with the FeNiMoO precursor in the process of phosphorization, resulting in the formation of the core-shell micropillar structure, as shown in Figure 1A. The core corresponds to NiMoO4 and the shell could be tentatively assigned to a hybrid containing MoO2 and NiFe phosphorous oxide.

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO<sub>4</sub> core-shell micropillar electrode

Figure 1. Scanning electron microscopy images of (A) a cross section of P-FeNiMoO and (B) P-NiFe@NiMoO4 electrode. (C-E) Transmission electron microscopy and high-resolution transmission electron microscopy images of P-NiFe@NiMoO4 scraped from nickel foam. (F) Elemental mapping images of P-NiFe@NiMoO4.

Figure 1B and Supplementary Figure 6 show the SEM images of the P-NiFe@NiMoO4 electrode at different magnifications, where it is clear that the surface of the micropillars became porous after the electrochemical oxidation process. The numerous pores offer large surface areas, active sites, efficient electrolyte diffusion and oxygen evolution channels. The A-FeNiMoO was electrochemically oxidized using exactly the same process as P-NiFe@NiMoO4 (denoted as A-NiFe@NiMoO4). The SEM images in Supplementary Figure 7 show that there is no apparent porous structure on the surface of A-NiFe@NiMoO4, which further confirms that the formation of the porous structure is closely connected with the etching of MoO2. The XRD pattern of P-NiFe@NiMoO4 in Supplementary Figure 5 shows the absence of diffraction peaks associated with MoO2 after the electrochemical oxidation process in comparison with that of P-FeNiMoO. It has been reported that low-valence MoO2 can be oxidized to high-valence MoO3, which could then be dissolved in an alkaline solution owing to the instability of MoO3[39,40]. Thus, the emergence of pores could be attributed to the etching of MoO2, which boosts the reconstruction of the shell during electrochemical oxidation[41], leading to the formation of the porous phosphate-intercalated NiFe (oxy)hydroxide shell.

To gain insights into the electrochemical oxidation process, in-situ Raman spectroscopy, as shown in Supplementary Figure 8, was conducted to investigate the surface reconstruction of the catalytic electrode with controlled potentials. For the pristine P-NiFe@NiMoO4 without any potential, the Raman bands at 704, 905 and 950 cm-1 can be assigned to the Mo-O vibrations in the MoO42- anions, while the bands at 196 and 344 cm-1 are assigned to the Mo-O vibrations in MoO2[42]. The Raman bands are stable and show a minor change when the potential is increased to 1.6 V. With the applied potential increased to 1.8 V, the MoO2 bands disappear while the NiFe (oxy)hydroxide bands appear at 465 and 545 cm-1, indicating the formation of NiFe (oxy)hydroxide under anodic potentials. The inner NiMoO4 remained stable even when the potential was increased to 2.2 V, in agreement with the literature[41]. It is evidenced that the surface reconstruction contributes to the porous structure.

Figure 1C and D and Supplementary Figure 9 show the TEM images. It is clear from the TEM image in Figure 1D that the pores, indicated by the orange ellipses, are present on the micropillar. It should be noted from the TEM image in Figure 1C that a hollow structure at the lower end of the micropillar, indicated by the orange ellipse, could be tentatively attributed to the fact that the significantly greater amount of PH3 gas produced during the decomposition of NaH2PO2 could react with the FeNiMoO precursor to generate a large amount of MoO2, the etching of which leads to the formation of the hollow structure instead of pores. The HRTEM image in Figure 1E reveals that the lattice distance of 0.233 nm could be attributed to the (002) plane of Ni(OH)2. The elemental mapping images and composition analysis of P-NiFe@NiMoO4 from energy-dispersive X-ray spectroscopy in Figure 1F and Supplementary Figure 10, respectively, confirm that the Fe, Ni, Mo, O and P elements are uniformly distributed.

The XPS surveys of the P-NiFe@NiMoO4 and P-FeNiMoO electrodes are shown in Figure 2A and Supplementary Figure 11, respectively, which further validate the existence of the Fe, Ni, Mo, O and P elements. Their contents are summarized in Supplementary Table 1. The decrease in Mo content from 7.23% to 0.36% proves the etching of MoO2 during the electrochemical oxidation process. It can be seen in Figure 2B that the two peaks located at 712.2 and 724.3 eV are assigned to the 2p3/2 and 2p1/2 levels of Fe(III), respectively, indicating the presence of the Fe3+ oxidation state[43]. The peak at 705.2 eV is assigned to Ni Auger electrons.

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO<sub>4</sub> core-shell micropillar electrode

Figure 2. (A) X-ray photoelectron spectroscopy survey and high-resolution spectra of (B) Fe 2p, (C) Ni 2p, (D) Mo 3d, (E) O 1s and (F) P 2p for P-NiFe@NiMoO4 electrode.

The high-resolution X-ray photoelectron spectrum of Ni 2p in Figure 2C demonstrates not only that the peaks located at 872.9 and 855.2 eV are from Ni(OH)2 with a binding energy gap of 17.7 eV[44,45] but also that the peaks located at 875.5 and 856.9 eV from NiMoO4 are observed[46]. In addition, the spectrum of Mo 3d in Figure 2D shows that the two peaks at 235.6 and 232.4 eV belong to Mo6+, thereby further confirming the presence of the inside core of NiMoO4 after the electrochemical oxidation process based on the presence of Ni2+ in Figure 2C[47]. The low valence of Mo is shown in Supplementary Figure 12A, which is consistent with the above TEM image and XRD results. As for the O 1s spectrum [Figure 2E], the three peaks at 532.9, 530.9 and 528.9 eV are assigned to H-O-H, metal-OH and metal-O, respectively[48]. The strong P-O peak at 134.3 eV in the P 2p spectrum [Figure 2F] is assigned to the presence of phosphate[49]. Nevertheless, the P-O peak at 133.5 eV in Supplementary Figure 12B negatively shifts, which indeed confirms the formation of NiFe phosphorous oxide after phosphorization. In summary, a porous phosphate-intercalated NiFe (oxy)hydroxide shell covered on NiMoO4, which consists of Ni(OH)2 and FeOOH, appears to have been formed through a surface reconstruction during the electrochemical oxidation process under high potentials.

To evaluate the OER performance of the P-NiFe@NiMoO4 catalysts, a three-electrode system was deployed in 1 M KOH. A-NiFe@NiMoO4, P-Ni@NiMoO4 and commercial RuO2 were also used as benchmarks for comparison. From the polarization curves shown in Figure 3A and Supplementary Figure 13, it can be seen that the P-NiFe@NiMoO4 electrode exhibits a significant improvement for the OER, compared with A-NiFe@NiMoO4 and P-Ni@NiMoO4 and commercial RuO2. Figure 3B delivers current densities of 100 and 200 mA/cm2 at small overpotentials of 238 and 257 mV in 1 M KOH, respectively. In addition, even at a higher current density of 300 mA/cm2, the required overpotential is 266 mV. The overpotential of the P-NiFe@NiMoO4 electrode is well below that of the other as-synthesized electrodes, such as A-NiFe@NiMoO4 and P-Ni@NiMoO4, despite the different current densities. The above results further confirm that the introduction of Fe and P plays a vital part in enhancing the OER performance. Notably, the P element is oxidized to phosphate, which intercalates into the NiFe (oxy)hydroxide layer to resist corrosion from Cl-.

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO<sub>4</sub> core-shell micropillar electrode

Figure 3. (A) Polarization curves and (B) comparison of overpotentials required to achieve current densities of 100, 200 and 300 mA cm-2 of P-NiFe@NiMoO4, A-NiFe@NiMoO4 and P-Ni@NiMoO4 electrodes and commercial RuO2 for the oxygen evolution reaction in 1 M KOH. Corresponding (C) Cdl values and (D) Tafel plots of the above electrodes. (E) Chronopotentiometric curve of P-NiFe@NiMoO4 electrode at a current density of 100 mA cm-2 in 1 M KOH.

The Tafel plot in Figure 3C shows that the P-NiFe@NiMoO4 electrode exhibits a smaller Tafel slope (54.3 mV/dec) compared with A-NiFe@NiMoO4 (83.1 mV/dec) and P-Ni@NiMoO4 (81.8 mV/dec), suggesting its rapid electrocatalytic kinetics. The high electrochemically active surface area is positively related to the number of active sites, which is calculated from the double-layer capacitance (Cdl). Figure 3D presents the Cdl values from the CV curves [Supplementary Figure 14]. The Cdl values of the P-NiFe@NiMoO4, A-NiFe@NiMoO4 and P-Ni@NiMoO4 electrodes are 9.2, 6.8 and 5.1 mF/cm2, respectively, implying the highly enhanced surface area and growing active sites implemented by forming the porous NiFe (oxy)hydroxide shell coated on the NiMoO4 micropillars. Electrochemical impedance spectroscopy [Supplementary Figure 15] shows that the charge-transfer resistance (Rct) of the P-NiFe@NiMoO4 electrode is minor (0.11 Ω) compared with A-NiFe@NiMoO4 (0.19 Ω) and P-Ni@NiMoO4 (0.51 Ω), thereby validating its outstanding electronic conductivity and charge transfer. Therefore, the improved OER activity benefits from the large surface area and efficient charge transfer of the porous shell on the NiMoO4 micropillars. Stability is another important parameter that is assessed by chronopotentiometric measurements of the P-NiFe@NiMoO4 electrode at a current density of 100 mA/cm2 in 1 M KOH at 25 °C. The chronopotentiometric curve of the P-NiFe@NiMoO4 electrode at a current density of 100 mA cm-2 in 1 M KOH in Figure 3E indicates that the electrochemical catalytic activity remains constant over 100 h without a noticeable decrease.

An efficient HER catalyst based on the P-NiMoO electrode [Supplementary Figure 16 and Supplementary Figure 17] was also successfully synthesized to combine with the P-NiFe@NiMoO4 electrode for overall seawater splitting. The HER performance of the P-NiMoO, P-FeNiMoO and A-NiMoO electrodes and commercial Pt/C were tested in 1 M KOH. Supplementary Figure 18A shows that the P-NiMoO electrode could deliver similar overpotentials to the P-FeNiMoO electrode below a current density of 50 mA/cm2 while exhibiting superior performance above 50 mA/cm2. The overpotential of the P-NiMoO electrode is only 149 mV at a current density of 100 mA/cm2, which is higher than that of the A-NiMoO and P-FeNiMoO electrodes. Supplementary Figure 18B and Supplementary Figure 18C show that the P-NiMoO electrode presents a smaller Tafel slope of 53 mV/dec and Rct of 0.7 Ω. The Cdl value [Supplementary Figure 18D] of the P-NiMoO electrode is 1.076 mF/cm2, calculated from the CV plots [Supplementary Figure 19]. Moreover, the stability test was carried out in 1 M KOH at a current density of 100 mA/cm2 and the result demonstrates no significant decrease over 100 h [Supplementary Figure 20].

The OER and HER activities of the P-NiFe@NiMoO4 and P-NiMoO electrodes were then evaluated in simulated seawater (1 M KOH + 0.5 M NaCl) and alkaline seawater (1 M KOH + seawater). The polarization curves in Figure 4A show that the P-NiFe@NiMoO4 electrode retains superior OER activity in 1 M KOH + 0.5 M NaCl, requiring overpotentials of 258, 290 and 312 mV to achieve 100, 200 and 300 mA/cm2, respectively [Figure 4B]. Moreover, in 1 M KOH + seawater, the activity decreases slightly owing to some insoluble precipitate[16,28]. Even in this situation, the overpotentials of the P-NiFe@NiMoO4 electrode, at current densities of 100, 200 and 300 mA/cm2, are still 258, 291 and 317 mV, respectively, which surpass those of most reported catalysts [Supplementary Table 2]. Furthermore, the HER activity of the P-NiMoO electrode was also tested in 1 M KOH + 0.5 M NaCl and 1 M KOH + seawater, which requires a similar overpotential of 146 mV at a current density of 100 mA/cm2 [Supplementary Figure 21].

Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO<sub>4</sub> core-shell micropillar electrode

Figure 4. (A) Polarization curves and (B) comparison of overpotentials at different current densities of 100, 200 and 300 mA cm-2 of P-NiFe@NiMoO4 electrode tested in different electrolytes. (C) Overall seawater spilling polarization curves in different electrolytes. Inset illustrates the overall two-electrode seawater splitting electrolyzer that paired P-NiFe@NiMoO4 with P-NiMoO. (D) Comparison between the amount of collected and theoretical gaseous products (H2 and O2) by the two-electrode electrolyzer at a constant current density of 40 mA cm-2 in 1 M KOH + seawater at 25 °C. (E) Durability tests of electrolyzer at a constant current density of 100 mA cm-2 in 1 M KOH + seawater at 25 °C. (F) Raman spectra of P-NiFe@NiMoO4.

A two-electrode electrolyzer was set up to further investigate the overall seawater splitting performance in different electrolytes [Figure 4C], where the P-NiFe@NiMoO4 electrode (anode) was coupled with the P-NiMoO electrode (cathode). Remarkably, this electrolyzer exhibits outstanding performance with a low cell voltage of ~1.63 V under a current density of 100 mA/cm2 in both the 1 M KOH + 0.5 M NaCl and 1 M KOH + seawater electrolytes. Even at larger current densities, the performance is still impressive, outperforming most previously reported electrolyzers [Supplementary Table 3]. The Faradaic efficiency was then evaluated in 1 M KOH + seawater by collecting the evolved H2 and O2 gases [Supplementary Figure 22]. As shown in Figure 4D, the molar ratio of H2 to O2 approaches 2:1 and the measured gas volume is in good agreement with the calculated volume, presenting high Faradaic efficiency and selectivity. Moreover, the as-synthesized P-NiFe@NiMoO4 || P-NiMoO electrolyzer presents excellent stability over 200 h in 1 M KOH + seawater [Figure 4E], while the stability of the A-NiFe@NiMoO4 || P-NiMoO electrolyzer decreases dramatically, indicating the promotion of stability in the presence of phosphate.

For further clarification of phosphate in the P-NiFe@NiMoO4 electrode, Raman spectroscopy [Figure 4F] was conducted, notably showing the presence of phosphate with peaks at 409.8, 588.2 and 948.8 cm-1 assigned to P-O vibrations[50]. The other two strong peaks, at 465.7 and 540.9 cm-1, are assigned to NiFe (oxy)hydroxide[17]. Peaks of MoO42- located at 706.8, 907.4 and 960.4 cm-1 can also be observed, which is consistent with the XRD result, confirming the NiMoO4 micropillars anchored on the NF substrate[39].

The morphology of the P-NiFe@NiMoO4 electrode after seawater splitting was investigated by SEM [Supplementary Figure 23]. After 200 h of electrolysis, the NiMoO4 micropillars still maintain their initial structure of being firmly anchored on the NF. On the surface of the NiMoO4 micropillars, nanoarrays were formed due to the reconstruction of the Ni and Fe elements during long-term stability under a high potential. Overall, the P-NiFe@NiMoO4 || P-NiMoO electrolyzer works well and delivers a superior cell voltage, indicating its potential application in hydrogen production.

CONCLUSIONS

In summary, we have successfully demonstrated the use of a porous anode made up of phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillars for active and stable seawater splitting. The introduction of Fe and the phosphorization are incredibly beneficial to the enhancement of activity and stability. During electrochemical oxidation, the etching of MoO2 results in a reconstruction process on the micropillars’ surface, which generates a porous phosphate-intercalated NiFe (oxy)hydroxide shell. The large pores endow the P-NiFe@NiMoO4 electrode with a large surface area, numerous active sites and efficient charge transfer. The P-NiFe@NiMoO4 electrode exhibits a low overpotential of 258 mV at a current density of 100 mA/cm2 in alkaline seawater. In addition, by pairing the anode with a P-NiMoO cathode, the two-electrode electrolyzer outputs a current density of 100 mA/cm2 at a low voltage of 1.63 V. The electrolyzer also exhibits superior stability over 200 h without any apparent activity loss, which is attributed to the passivating anions of phosphate against chloride corrosion. This work provides insight into the development of the industrial application of seawater splitting for green hydrogen production.

DECLARATIONS

Authors’ contributions

Methodology, investigation, formal analysis, writing manuscript, visualization: Yang C, Gao N

Validation, resources, formal analysis: Wang X

Validation, investigation: Lu J, Li Y

Formal analysis, validation: Cao L

Project administration, conceptualization, funding acquisition, supervision, writing manuscript: Liang HP

Availability of data and materials

The data supporting our findings can be found in the supplementary information.

Financial support and sponsorship

This work was supported by the DICP and QIBEBT (Grant UN201804), Dalian National Laboratory for Clean Energy (DNL), CAS and Research Innovation Fund (Grant QIBEBTSZ201801), the Qingdao Leading Talents of Innovation and Entrepreneurship (No. 19-3-2-11-2hc), Innovation Academy for Green Manufacture (Grant IAGM2020C14), the DNL Cooperation Fund, CAS (DNL202010), China Postdoctoral Science Foundation funded project (Grant No. 2020M682250) and the Natural Science Foundation of Shandong (Grant ZR202102220330).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2021.

Supplementary Materials

Supplementary Materials

REFERENCES

1. Cox CR, Lee JZ, Nocera DG, Buonassisi T. Ten-percent solar-to-fuel conversion with nonprecious materials. Proc Natl Acad Sci U S A 2014;111:14057-61.

2. Faber MS, Jin S. Earth-abundant inorganic electrocatalysts and their nanostructures for energy conversion applications. Energy Environ Sci 2014;7:3519-42.

3. Wang X, Yang C, Wang X, et al. Green synthesis of a highly efficient and stable single-atom iron catalyst anchored on nitrogen-doped carbon nanorods for the oxygen reduction reaction. ACS Sustainable Chem Eng 2021;9:137-46.

4. Tiwari JN, Sultan S, Myung CW, et al. Multicomponent electrocatalyst with ultralow Pt loading and high hydrogen evolution activity. Nat Energy 2018;3:773-82.

5. Turner JA. Sustainable hydrogen production. Science 2004;305:972-4.

6. Cabán-Acevedo M, Stone ML, Schmidt JR, et al. Efficient hydrogen evolution catalysis using ternary pyrite-type cobalt phosphosulphide. Nat Mater 2015;14:1245-51.

7. Luo J, Im JH, Mayer MT, et al. Water photolysis at 12.3% efficiency via perovskite photovoltaics and Earth-abundant catalysts. Science 2014;345:1593-6.

8. Pu Z, Zhao J, Amiinu IS, et al. A universal synthesis strategy for P-rich noble metal diphosphide-based electrocatalysts for the hydrogen evolution reaction. Energy Environ Sci 2019;12:952-7.

9. Zhang J, Zhang Q, Feng X. Support and interface effects in water-splitting electrocatalysts. Adv Mater 2019;31:e1808167.

10. Seh ZW, Kibsgaard J, Dickens CF, Chorkendorff I, Nørskov JK, Jaramillo TF. Combining theory and experiment in electrocatalysis: insights into materials design. Science 2017;355:eaad4998.

11. Du J, Li C, Wang X, Shi X, Liang HP. Electrochemical synthesis of cation vacancy-enriched ultrathin bimetallic oxyhydroxide nanoplatelets for enhanced water oxidation. ACS Appl Mater Interfaces 2019;11:25958-66.

12. Li W, Li F, Yang H, et al. A bio-inspired coordination polymer as outstanding water oxidation catalyst via second coordination sphere engineering. Nat Commun 2019;10:5074.

13. Shi Q, Zhu C, Du D, Lin Y. Robust noble metal-based electrocatalysts for oxygen evolution reaction. Chem Soc Rev 2019;48:3181-92.

14. Yu L, Zhou H, Sun J, et al. Amorphous NiFe layered double hydroxide nanosheets decorated on 3D nickel phosphide nanoarrays: a hierarchical core-shell electrocatalyst for efficient oxygen evolution. J Mater Chem A 2018;6:13619-23.

15. Yu L, Zhu Q, Song S, et al. Non-noble metal-nitride based electrocatalysts for high-performance alkaline seawater electrolysis. Nat Commun 2019;10:5106.

16. Tong W, Forster M, Dionigi F, et al. Electrolysis of low-grade and saline surface water. Nat Energy 2020;5:367-77.

17. Kuang Y, Kenney MJ, Meng Y, et al. Solar-driven, highly sustained splitting of seawater into hydrogen and oxygen fuels. Proc Natl Acad Sci U S A 2019;116:6624-9.

18. Lv F, Zhang W, Yang W, et al. Ir-based alloy nanoflowers with optimized hydrogen binding energy as bifunctional electrocatalysts for overall water splitting. Small Methods 2020;4:1900129.

19. Dong J, Duan L, Wu Q, Yao W. Facile preparation of Pt/CdS photocatalyst by a modified photoreduction method with efficient hydrogen production. Int J Hydrogen Energy 2018;43:2139-47.

20. Liu S, Huang S. The role of interface charge transfer on Pt based catalysts for water splitting. Int J Hydrogen Energy 2018;43:15225-33.

21. Badiei YM, Xie Y, Renderos G, et al. Rapid identification of homogeneous O2 evolution catalysts and comparative studies of Ru(II)-carboxamides vs. Ru(II)-carboxylates in water-oxidation. J Catal 2019;369:10-20.

22. Duan Y, Yu Z, Hu S, et al. Scaled-up synthesis of amorphous NiFeMo oxides and their rapid surface reconstruction for superior oxygen evolution catalysis. Angew Chem 2019;131:15919-24.

23. Wu T, Sun S, Song J, et al. Iron-facilitated dynamic active-site generation on spinel CoAl2O4 with self-termination of surface reconstruction for water oxidation. Nat Catal 2019;2:763-72.

24. Du J, Li C, Wang X, Jones TG, Liang H. Cobalt oxyhydroxide with highly porous structures as active and stable phase for efficient water oxidation. Electrochimica Acta 2019;303:231-8.

25. Chen G, Du J, Wang X, Shi X, Wang Z, Liang H. Iron-induced 3D nanoporous iron-cobalt oxyhydroxide on carbon cloth as a highly efficient electrode for oxygen evolution reaction. Chinese J Catal 2019;40:1540-7.

26. Huang Z, Song J, Du Y, et al. Chemical and structural origin of lattice oxygen oxidation in Co-Zn oxyhydroxide oxygen evolution electrocatalysts. Nat Energy 2019;4:329-38.

27. Lv Q, Han J, Tan X, Wang W, Cao L, Dong B. Featherlike NiCoP holey nanoarrys for efficient and stable seawater splitting. ACS Appl Energy Mater 2019;2:3910-7.

28. Wang S, Yang P, Sun X, et al. Synthesis of 3D heterostructure Co-doped Fe2P electrocatalyst for overall seawater electrolysis. Applied Catalysis B: Environmental 2021;297:120386.

29. Wu L, Yu L, Zhang F, et al. Heterogeneous bimetallic phosphide Ni2P-Fe2P as an efficient bifunctional catalyst for water/seawater splitting. Adv Funct Mater 2021;31:2006484.

30. Jin H, Liu X, Vasileff A, et al. Single-crystal nitrogen-rich two-dimensional Mo5N6 nanosheets for efficient and stable seawater splitting. ACS Nano 2018;12:12761-9.

31. Wang B, Lu M, Chen D, et al. NixFeyN@C microsheet arrays on Ni foam as an efficient and durable electrocatalyst for electrolytic splitting of alkaline seawater. J Mater Chem A 2021;9:13562-9.

32. Gupta S, Forster M, Yadav A, Cowan AJ, Patel N, Patel M. Highly efficient and selective metal oxy-boride electrocatalysts for oxygen evolution from alkali and saline solutions. ACS Appl Energy Mater 2020;3:7619-28.

33. Li Y, Wu X, Wang J, et al. Sandwich structured Ni3S2-MoS2-Ni3S2@Ni foam electrode as a stable bifunctional electrocatalyst for highly sustained overall seawater splitting. Electrochimica Acta 2021;390:138833.

34. Wang C, Zhu M, Cao Z, et al. Heterogeneous bimetallic sulfides based seawater electrolysis towards stable industrial-level large current density. Applied Catalysis B: Environmental 2021;291:120071.

35. Chang J, Wang G, Yang Z, et al. Dual-doping and synergism toward high-performance seawater electrolysis. Adv Mater 2021;33:e2101425.

36. Hung W, Xue B, Lin T, Lu S, Tsao I. A highly active selenized nickel-iron electrode with layered double hydroxide for electrocatalytic water splitting in saline electrolyte. Mater Today Energy 2021;19:100575.

37. Dionigi F, Reier T, Pawolek Z, Gliech M, Strasser P. Design criteria, operating conditions, and nickel-iron hydroxide catalyst materials for selective seawater electrolysis. ChemSusChem 2016;9:962-72.

38. Liu J, Ding P, Zhu Z, et al. Engineering self-reconstruction via flexible components in layered double hydroxides for superior-evolving performance. Small 2021;17:e2101671.

39. Wang W, Yang Z, Jiao F, Gong Y. (P, W)-codoped MoO2 nanoflowers on nickel foam as an efficient bifunctional electrocatalyst for overall water splitting. Appl Surf Sci 2020;529:146987.

40. Xu C, Lu W, Yan L, et al. Hierarchical molybdenum-doped cobaltous hydroxide nanotubes assembled by cross-linked porous nanosheets with efficient electronic modulation toward overall water splitting. J Colloid Interface Sci 2020;562:400-8.

41. Solomon G, Landström A, Mazzaro R, et al. NiMoO4@Co3O4 core-shell nanorods: in situ catalyst reconstruction toward high efficiency oxygen evolution reaction. Adv Energy Mater 2021;11:2101324.

42. Zhang C, Lu J, Li M, et al. Transfer hydrogenation of nitroarenes with hydrazine at near-room temperature catalysed by a MoO2 catalyst. Green Chem 2016;18:2435-42.

43. Yu L, Wu L, Mcelhenny B, et al. Ultrafast room-temperature synthesis of porous S-doped Ni/Fe (oxy)hydroxide electrodes for oxygen evolution catalysis in seawater splitting. Energy Environ Sci 2020;13:3439-46.

44. Xiong K, Huang L, Gao Y, et al. Formation of a thin-layer of nickel hydroxide on nickel phosphide nanopillars for hydrogen evolution. Electrochem commun 2018;92:9-13.

45. Wu J, Subramaniam J, Liu Y, Geng D, Meng X. Facile assembly of Ni(OH)2 nanosheets on nitrogen-doped carbon nanotubes network as high-performance electrocatalyst for oxygen evolution reaction. J Alloys Compd 2018;731:766-73.

46. Zhuang S, Tong S, Wang H, et al. The P/NiFe doped NiMoO4 micro-pillars arrays for highly active and durable hydrogen/oxygen evolution reaction towards overall water splitting. Int J Hydrogen Energy 2019;44:24546-58.

47. Li YK, Zhang G, Lu WT, Cao FF. Amorphous Ni-Fe-Mo suboxides coupled with Ni network as porous nanoplate array on nickel foam: a highly efficient and durable bifunctional electrode for overall water splitting. Adv Sci (Weinh) 2020;7:1902034.

48. Zhang H, Li X, Hähnel A, et al. Bifunctional heterostructure assembly of NiFe LDH nanosheets on NiCoP nanowires for highly efficient and stable overall water splitting. Adv Funct Mater 2018;28:1706847.

49. Wang F, Ma K, Tian W, et al. P-Doped NiMoO4 parallel arrays anchored on cobalt carbonate hydroxide with oxygen vacancies and mass transfer channels for supercapacitors and oxygen evolution. J Mater Chem A 2019;7:19589-96.

50. Anjaneyulu U, Pattanayak DK, Vijayalakshmi U. Snail shell derived natural hydroxyapatite: effects on NIH-3T3 cells for orthopedic applications. Mater Manuf Process 2015;31:206-16.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Yang C, Gao N, Wang X, Lu J, Cao L, Li Y, Liang HP. Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillar electrode. Energy Mater 2021;1:100015. http://dx.doi.org/10.20517/energymater.2021.16

AMA Style

Yang C, Gao N, Wang X, Lu J, Cao L, Li Y, Liang HP. Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillar electrode. Energy Materials. 2021; 1(2): 100015. http://dx.doi.org/10.20517/energymater.2021.16

Chicago/Turabian Style

Yang, Chen, Nannan Gao, Xilong Wang, Jiajia Lu, Lijuan Cao, Yadong Li, Han-Pu Liang. 2021. "Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillar electrode" Energy Materials. 1, no.2: 100015. http://dx.doi.org/10.20517/energymater.2021.16

ACS Style

Yang, C.; Gao N.; Wang X.; Lu J.; Cao L.; Li Y.; Liang H.P. Stable and efficient seawater splitting on a porous phosphate-intercalated NiFe (oxy)hydroxide@NiMoO4 core-shell micropillar electrode. Energy Mater. 2021, 1, 100015. http://dx.doi.org/10.20517/energymater.2021.16

About This Article

© The Author(s) 2021. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
26819
Downloads
3657
Citations
Comments
0
13

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 26 clicks
Like This Article 13 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/