Gwaza Eric
Ayom
a,
Malik Dilshad
Khan
ab,
Jonghyun
Choi
c,
Ram Krishna
Gupta
c,
Werner E.
van Zyl
d and
Neerish
Revaprasadu
*a
aDepartment of Chemistry, University of Zululand, Private Bag X1001, KwaDlangezwa 3880, South Africa. E-mail: RevaprasaduN@unizulu.ac.za
bInstitute of Physical Chemistry, Polish Academy of Sciences, Kasprzaka 44/52, 01-224, Warsaw, Poland
cDepartment of Chemistry, Pittsburg State University, Pittsburg, KS 66762, USA
dSchool of Chemistry and Physics, University of KwaZuluNatal, Westville Campus, Chiltern Hills, Private Bag X54001, Durban, 4000, South Africa
First published on 31st July 2021
Cost-effective and readily available catalysts applicable for electrochemical conversion technologies are highly desired. Herein, we report the synthesis of dithiophosphonate complexes of the type [Ni{S2P(OH)(4-CH3OC6H4)}2] (1), [Co{S2P(OC4H9)(4-CH3OC6H4)}3] (2) and [Fe{S2P(OH)(4-CH3OC6H4)}3] (3) and employed them to prepare Ni2P, Co-Ni2P and Fe-Ni2P nanoparticles. Ni2P was formed by a facile hot injection method by decomposing complex 1 in tri-octylphosphine oxide/tri-n-octylphosphine at 300 °C. The prepared Ni2P was doped with Co and Fe employing complexes 2 and 3, respectively, under similar experimental conditions. Doping Ni2P with Co and Fe demonstrated synergistic improvement of Ni2P performance as an electrocatalyst in supercapcitance, hydrogen evololution and oxygen evolution reactions in alkaline medium. Cobalt doping improved the Ni2P charge storage capacity with a supercapacitance of 864 F g−1 at 1 A g−1 current density. Fe doped Ni2P recorded the lowest overpotential of 259 mV to achieve a current density of 10 mA cm−2 and a Tafel slope of 80 mV dec−1 for OER, better than the undoped Ni2P and the benchmark IrO2. Likewise, Fe-doped Ni2P electrode required the lowest overpotential of 68 mV with a Tafel slope of 110 mV dec−1 to attain the same current density for HER. All catalysts showed excellent stability in supercapacitance and overall water splitting reactions, indicating their practical use in energy conversion technologies.
Nickel phosphides continue to attract attention as catalysts in water splitting and supercapacitance applications.6–8 The growing attention is due to their abundant earth reserves, high electrical conductivity, fast charge transfer, and good reaction kinetics.9 Nickel phosphide is a binary system with multiple compositions or phases depending on the ratio of nickel to phosphorus in the system. Its different compositions range from metal-rich phases such as Ni2P, Ni12P5, Ni3P, Ni7P3 and Ni5P4 to phosphide rich ones like NiP2 and NiP3.7,10
Besides tuning phase, shape and size, doping of different metal atoms into the metal phosphides is another facile and useful strategy employed by researchers to improve the performance of these materials as catalysts for supercapacitors and water splitting.11 Incorporating a foreign atom in metal phosphides has been illustrated by density functional theory as a route to modifying the adsorption/desorption energies of reactants/products, which is crucial for the activity of a catalyst,7,12 as well as modulating the density states at the Fermi level and hence improving the reaction kinetics.13 Moreover, the introduction of another metal in a metal phosphide crystal lattice can cause the redistribution of valence electrons, which in turn provide two electron donation sites leading to improved catalyst performance.14 For example, Liu et al. doped Ni2P with nitrogen to improve the performance of their electrodes in supercapacitance examinations.15 Lin et al. recently fabricated an iron and oxygen co-doped nickel phosphide that showed good activity in overall water splitting reactions.16 The performance of nickel phosphide, an excellent catalyst for water splitting7 can be improved by incorporating Co into its lattice.13 For example, Qiu et al. showed by density functional theory (DFT) analysis that incorporating Co in nickel phosphide leads to an increase in active sites for OH− adsorption that improves OER activity.43 Co and Fe-based electrodes have high calculated theoretical supercapacitance values.1717,43 The incorporation of a foreign atom into metal phosphides has therefore been demonstrated as a strategy to improve the electronic structure, transfer capability, and density active sites for improved catalytic performance.11 A survey of the literature, however, shows gaps in doping as a strategy to improve Ni2P performance (especially Co and Fe) in supercapacitors and water splitting.
Synthesis of phase pure nickel phosphide is challenging due to the existence of nickel phosphide in different compositions. The use of molecular compounds as single-source precursors for the preparation of nanoparticles is well documented.18 The strengths of this route in nanoparticle preparation over the use of multiple sources lie in the control of ligand design and hence the formation of the suitable metal-phosphide/chalcogenide bonds suitable for the preparation of desired nanomaterials.5 The exploration of single-source molecular precursors to prepare nickel phosphides is, however, scarce.7,19 Lukehart and Milne employed a nickel phosphine complex, tetrakis{diphenyl[2-(triethoxysilyl)-ethyl]phosphine}nickel(0), to prepare nickel phosphide nanoclusters in the first report of the use of single-source precursors to make nickel phosphides.20 Maneeprakorn et al. in another study employed dithiophosphinates ([Ni(Se2PiPr2)2], [Ni(Se2PtBu2)2] and [Ni(Se2PPh2)2]) as molecular precursors to prepare Ni2P and Ni5P4.19 Pan et al. formed the bis(triphenylphosphine)nickel dichloride complex from the reaction of triphenylphosphine and nickel chloride hexahydrate which was exploited to form nickel phosphides. Habas et al. similarly employed triphenylphosphine and a commercially available air-stable nickel phosphine complex [Ni(PPh3)2(CO)2] to prepare dinickel phosphide.21 We also recently demonstrated the solvent-less synthesis of pure phase dinickel phosphide employing dithiophosphonate complexes and triphenylphosphine.8
The scarcity of nickel phosphides in the literature particularly formed via the single-source precursor route compared to the sulfides or oxides, is attributed to the synthetic difficulties in the formation of desired molecular complexes. Dithiophosphonate ligands, which are phosphorus and sulfur-containing, are by far the less studied class of the phosphor-1,1-dithiolates. They have rarely been employed to prepare phosphides. Likewise, the preparations of nickel phosphides are constrained by their prolonged reaction time, elevated temperatures9,22 and the toxicity of reagents usually employed.23 Furthermore, using precursors with similar structural features for dopants may result in easy and efficient incorporation of dopants, as all the precursors will follow a similar kind of decomposition route.
Herein, we have prepared three dithiophosphonate complexes; [Ni{S2P(OH)(4-CH3OC6H4)}2] (1), [Co{S2P(OC4H9)(4-CH3OC6H4)}3] (2) and [Fe{S2P(OH)(4-CH3OC6H4)}3] (3) and explored them as molecular precursors to prepare Ni2P and Co and Fe-doped Ni2P with tri-octylphosphine (TOP) as the phosphorus source via the hot injection method. The electrocatalytic performance of nickel phosphide and the effect of different concentrations of Co and Fe doping on the performance of the prepared Ni2P was also investigated for overall water splitting and supercapacitance.
Complexes 1, 2 and 3 were confirmed by elemental analysis.
[Ni{S2P(OH)(4-CH3OC6H4)}2] (1): Calc. C: 33.82%; H: 3.24%; S: 25.79%. Found. C: 33.78%; H: 3.30%; S: 25.65%.
[Co{S2P(OC4H9)(4-CH3OC6H4)}3] (2): Calc. C: 44.79%; H: 5.47%; S: 21.74%. Found. C: 44.50%; H: 5.42%; S: 21.69%.
[Fe{S2P(OH)(4-CH3OC6H4)}3] (3): C: 44.95%; H: 5.49%; S: 21.81%. Found. C: 45.0%; H: 5.41%; S: 21.79%.
Electrochemical characterization of the prepared particles was performed using Gamry Potentiostat, which employed a three-electrode system. Sample preparation for electrochemical examinations was done by forming pastes of the particles using the particles (80 wt%), polyvinylidene difluoride (PVDF, 10 wt%), and acetylene black (10 wt%), which was prepared using N-methyl pyrrolidinone (NMP) as a solvent. These formed pastes were then applied to pre-cleaned and weighted nickel foams. Dried pastes were then employed as working electrodes. Ni foams (MTI Corporation, USA), 99.99% purity, were used for these investigations. Commercial carbons were employed as conducting acetylene black (MTI Corporation, USA) with particle sizes ranging between 35–40 nm. Platinum wires and saturated calomel electrodes (SCE) were utilized as counter and reference electrodes, respectively. All the examinations for supercapacitance and electrocatalysis were carried out using 3 M and 1 M KOH electrolyte, respectively. Charge storage capacity was measured using cyclic voltammetry (CV) and galvanostatic charge–discharge (CD) at different scan rates and current densities. Electrocatalytic properties of the prepared electrodes were studied via cyclic voltammetry and linear sweep voltammetry (LSV), with LSV done at a scan rate of 2 mV s−1 for OER measurements. Electrochemical impedance spectroscopic (EIS) was performed in the frequency range of 0.05 Hz to 10 kHz at an applied AC amplitude of 10 mV.
The hot injection decomposition of complex 1 employing trioctylphosphine oxide (TOPO) as a surfactant and trioctylphosphine (TOP) as a dispersing solvent for 1 hour yielded nickel phosphide, which is consistent with our previous study.5 The formed nickel phosphide matched well to hexagonal Ni2P (ICDD# 01-089-2742). It is interesting to note that TOP has been shown to be the phosphorus source in this formation of nickel phosphide despite the presence of intramolecular phosphorus atoms in complex 1.5 Doping Ni2P with 5 or 10% of complexes 2 (Co) and 3 (Fe) via the hot injection route did not change the phase of the nickel phosphide obtained or introduce any impurity. The phase purity of all nanomaterials prepared and the effect of doping was evaluated by p-XRD and is shown in Fig. 1a. The sharp p-XRD diffraction patterns of Ni2P, Co and Fe-doped Ni2P are well matched to pure hexagonal Ni2P and are indexed as per standard reference pattern (ICDD# 01-089-2742), indicating high crystallinity. The absence of any Ni, NiS or NiO diffraction peaks in the doped Ni2P indicates phase purity and successful incorporation of the dopants in Ni2P crystal structure. In all the samples, diffraction from the (111) plane was found to be more intense indicative of the significant growth direction of the hexagonal Ni2P. A closer examination of the most intense peak at ≈2θ = 41 (Fig. 1b) shows a slight shift to a lower angle for both increased Co and Fe-doping at the (111) plane of Ni2P. This could be due to the substitution of Ni2+ ions by the larger Co2+ or Fe2+ ions, as supported by literature.26,27
The introduction of dopants could result in the modification of the electronic and optical properties of a material. The optical properties of the nanoparticles were therefore analyzed, and the spectra are shown in Fig. S2.† All the particles absorb photons within the UV–vis region of between 250 and 300 nm. The absorption peaks of the doped Ni2P particles are slightly red-shifted relative to that of Ni2P, which could be indicative of the Co and Fe dopants acting as auxochromes.
Scanning electron microscopy (SEM) images were used to observe the morphology of the synthesized nanoparticles. The SEM images reveal highly agglomerated spherically clustered compact particles that are well distributed, with the compactness of the nanoparticles decreasing as the Ni2P nanoparticles are doped with Co and Fe (Fig. 2 and S3†). For a better insight into the morphology, transmission electron microscopy (TEM) was used, and the images are shown in Fig. S4.† The particles showed large irregular sheet-like morphology with smaller particles anchored/stacked on them. We also employed selected area electron diffraction (SAED) and high-resolution transmission electron microscopy (HRTEM) to further investigate the crystallinity and the micro-structure of as-synthesized dinickel phosphides (Fig. 2 and S4†). SAED images showed some well-defined spots indicative of crystallinity. However, the surface capping of catalysts with long alkyl chain surfactants may also result in reduced crystallinity to some extent. HRTEM images further confirmed the sheet-like morphology and crystallinity by showing clear lattice fringes that indexed well with the d-spacing from standard reference patterns for hexagonal Ni2P.
The elemental composition of the prepared materials was undertaken to examine the composition of Ni2P and the doped Ni2P. The energy-dispersive X-ray spectroscopy (EDX) emissions of all prepared materials showed only Ni and P for Ni2P and Co or Fe dopants for the doped Ni2P, which is in agreement with the powder XRD. EDX elemental mapping of all the samples confirms the good distribution of the elements as well as the doping of Co and Fe into Ni2P crystal lattice (Fig. 2). X-ray photoelectron spectroscopy (XPS) analyses of undoped Ni2P and doped Ni2P (Fe-Ni2P and Co-Ni2P) was performed since it can provide a direct evidence for the introduction of dopants that modulate the electronic structure of catalysts28,29 and is given in Fig. 3 and S5.† The survey spectra of the Ni2P, (Fe-Ni2P and Co-Ni2P) samples show the presence of Ni, P, O, C, (Fe and Co) elements as expected. The presence of oxygen and carbon in the samples is attributed to surface oxidation8 and adsorption.30 Moreover the intense peak for carbon is due to the presence of capping agents on the surface of nanomaterials. Survey spectra analysis of these samples indicates the incorporation of Fe and Co in the doped Ni2P which is in agreement with the p-XRD and SEM-EDX results. The XPS peak at 856.7 eV of the Fe-Ni2P is assigned to Ni2+ in the Ni 2p3/2 window while the satellite peak at 862.7 eV in the same window is attributed to surface oxidation (Fig. 3).8 The peaks at 874.2, 875.9 and 880.4 eV are assigned to Ni of the Fe-Ni2P in the Ni 2p1/2 window. Similar results were observed for the Ni2P and the Co-Ni2P samples as shown in Fig. S5.† The P 2p XPS regions (Fig. 3 and S5†) of these samples show peaks at 133, 134.6 and 130.9 eV for Ni2P, Fe-Ni2P and Co-Ni2P in the P 2p3/2 window, respectively, which are typical of the P of metal phosphides.30 The Co-Ni2P material has another peak at 127.3 eV in the P 2p ½ window assigned to POx species which could be attributed to oxidation of the phosphorus in the sample.30 High resolution spectra of the Fe 2p region of Fe-Ni2P (Fig. 3) indicated two peaks at 713 and 725.8 eV in the Fe 2p3/2 and Fe 2p1/2 windows, respectively, assigned to Fe3+.30 Similarly the Co 2p window of the XPS results of the Co-Ni2P show two peaks at 783.1 and 800.6 eV in the Co 2p3/2 and Co 2p1/2 windows, respectively, which may indicate presence of Co2+ and Co3+ as co-existing species.28 These XPS results confirm surface capping of nanoparticles, surface oxidation and also show the successful doping of Fe and Co into Ni2P.
Ni2+ + 2OH− → Ni(OH)2 | (1) |
Ni(OH)2 + OH− → NiOOH + H2O + e− | (2) |
We observed that the area under the CV curves of all the electrodes increases with an increase in scan rate, which is attributed to diffusion-limited kinetics.31Eqn (3) was employed to gain further insight into the diffusion-limited kinetics of the electrodes.34
i = avb | (3) |
In eqn (3), i is the peak current, v is the scan rate (mV s−1), while a and b are variable parameters. The charge storage mechanism of electrodes can be separated by capacitive and diffusion-limited contributions. The charge storage mechanism of an electrode is therefore based on parameter b in eqn (3). The charge storage mechanism could be diffusion-limited when b = 0.5 or capacitive when b = 1. The b values of 0.51, 0.54, and 0.51 were obtained for Ni2P, Co-Ni2P, and Fe-Ni2P electrodes, respectively, suggesting that diffusion-limited faradaic reaction dominates over the capacitive process in contribution to the charge storage capacity for all the electrodes (Fig. 5a). The dominance of diffusion-limited faradiac process suggests that the synthesized materials behave more like a battery, rather than the capacitance materials. To estimate the diffusion-controlled and capacitive contributions to the total energy storage of all the electrodes, a low and high scan rate of 10 and 100 mV s−1 was evaluated employing eqn (4),34
i = k1v + k2v1/2 | (4) |
Fig. 6a–c and S6† give the galvanostatic charge–discharge graphs of the five electrodes indicating the potential vs. time charge–discharge characteristics of the electrodes at different current densities. As shown in all the graphs, the plateau region was observed after a sharp potential drop. These plateau regions are attributed to the faradaic redox reactions, which corresponds to the CV results.32 The Ni2P, 5% Co-Ni2P, and 5% Fe-Ni2P electrodes recorded the specific capacitance of 674, 864, 856 F g−1 at 1A g−1 current density, respectively, based on the galvanostatic charge–discharge profiles. The recorded specific capacitance shows that doping Ni2P with Co and Fe improves its supercapacitance properties. The superior performance of the Co-doped Ni2P electrode compared to that of the Fe-doped Ni2P electrode may be due to the superior theoretical specific capacitance of Co-based electrodes compared to that of Fe-based ones. Lu et al. employed an organic-phase strategy to synthesize Ni2P electrodes that recorded a specific capacitance of 418 F, which is 40% lower than our Ni2P electrode (674 F g−1) at the same 1A g−1 current density.33 The synthesized Ni2P electrodes in that study were fabricated into a composite by coating with Ni, which improved their specific capacitance to 581 F g−1 compared to ours, where the doped Ni2P electrodes had 864 F g−1 (Co-Ni2P) and 856 F g−1 (Fe-Ni2P), respectively. Liu et al. recently prepared a series of Ni2P doped polypyrrole composites and tested their electrochemical application as supercapacitors.35 Their 30% Ni2P doped polypyrrole composite had the best performance of 476.5 F g−1, which is 29% lower than our prepared Ni2P (674 F g−1), 45% lower than Co-Ni2P (864 F g−1), and 44% lower than our Fe-Ni2P (856 F g−1) electrodes, respectively at 1A g−1 current density. A comparison of the supercapacitance of our electrodes to other similar nickel phosphide materials shows that our materials outperform other electrodes under similar conditions (Table S1†). The specific capacitance of an electrode generally decreases progressively with current density due to increasing limitation or infiltration of electrons and ions into the electrode surface. We, consequently, plotted the calculated specific capacitance values as a function of the current density, which is given in Fig. 6d and S6.† The specific capacitance of 417, 554, and 530 F g−1 was recorded for our Ni2P, Co-Ni2P, and Fe-Ni2P electrodes at a high current density of 30 A g−1. This result does not only indicate excellent stability of charge storage capacities for our electrodes but also shows that doping Ni2P with Co and Fe improves the Ni2P electrodes’ charge retention capacity.
![]() | ||
Fig. 6 GCD curves of (a) Ni2P, (b) Co-Ni2P, (c) Fe-Ni2P electrodes at various current densities and (d) variation of specific capacitance as a function of current density. |
We also measured the long-term stability of the Ni2P, Co-Ni2P, and Fe-Ni2P electrodes in 3000 charge–discharge cycles, as shown in Fig. 7. The electrodes showed a high retention capacity of 89, 85 and 74% of the initial cycles after 3000 cycles, along with 99% coulombic efficiency for Ni2P, Co-Ni2P and Fe-Ni2P electrodes, respectively, suggesting good cyclic stability. The CV and GCD measurements show that doping our Ni2P electrodes with Co and Fe dramatically improved its energy storage properties.
![]() | ||
Fig. 7 Capacitance retention and coulombic efficiency of (a) Ni2P, (b) Co-Ni2P, (c) Fe-Ni2P for 3000 cycles. |
![]() | ||
Fig. 8 (a) Polarization curves, (b) Tafel slopes of all samples for OER process, and (c) Nyquist plot at 0.6 V (V vs. SCE). |
Table S2† gives a further comparison of our electrodes as OER catalysts with respect to the OER activity of some state-of-the-art nickel phosphide-based electrocatalysts. The electrochemical impedance spectroscopy (EIS) measurements were employed to further study the effect of doping Co and Fe into the Ni2P electrode. These measurements (Nyquist plots), which are typically used to explore the charge transfer processes at the electrode, were obtained within the frequency range of 0.05 Hz to 10 kHz with applied AC amplitude of 10 mV at 0.6 V (V vs. SCE) and are given in Fig. 8c and S7.† In the obtained Nyquist plots, two arcs are seen in Ni2P, 5% Co-Ni2P, Co-Ni2P and 5% Fe-Ni2P electrodes and only one semicircle was observed in the Fe-Ni2P electrode. One arc at the high frequency represents the charge transfer resistance; the other arc at the low frequency represents the mass transfer resistance.40 The diameter of the arc or semicircle is a function of the charge transfer resistance.36 The smaller diameter of the arc at the high frequency for two of our electrodes indicates less charge transfer resistance and thus better charge transfer in the order Fe-Ni2P > Co-Ni2P > Ni2P. The 5% Co-Ni2P electrode had less charge transfer resistance compared to the 5% Fe-Ni2P though both of them did not improve the charge transfer resistance of the Ni2P electrode. In the case of arcs at low frequency, there is none for the Fe-Ni2P graph but one in 5% Co-Ni2P, Co-Ni2P, 5% Fe-Ni2P and Ni2P graphs. The smaller diameter of the arc at low frequency for 5% Co-Ni2P, Co-Ni2P and 5% Fe-Ni2P compared to Ni2P indicate that these electrodes have less mass-transfer resistance than the Ni2P electrode. These results of charge transfer resistance put together show that doping our Ni2P with Co and Fe generally improves its OER electrocatalytic activity.
The long-term stability of a catalyst is crucial to its practical applications. The durability of our electrodes was therefore investigated employing the chronoamperometry (CA) test and polarization curve measurements, which are shown in Fig. 9a–d and S7.† CA measurements were done at 0.55 V (V vs. SCE), and a stable current density was delivered for all samples over 35 hours. A slight fluctuation was observed in the graphs for all samples during the CA test due to the bubbling caused by the occurrence of oxygen gas during oxygen evolution.41 A measurement of the polarization curves for the 1st cycle and the 1000 cycles showed a negligible deviation between two curves for all the electrodes (Fig. 9b–d and S7†), indicating the excellent stability or durability of our electrodes as OER catalysts.
![]() | ||
Fig. 9 (a) Chronoamperometry (CA) at 0.55 V (V vs. SCE) for all samples. OER polarization curves at different cycles for (b) Ni2P, (c) Co-Ni2P, and (d) Fe-Ni2P. |
We also examined the electrocatalytic activity of our electrodes for hydrogen evolution reactions. Fig. 10 and S8† give the polarization curves and Tafel slopes of all the samples for the HER. The Ni2P, 5% Co-Ni2P, Co-Ni2P, 5% Fe-Ni2P and Fe-Ni2P electrodes recorded an overpotential of 164, 156, 158, 68 and 202 mV to achieve a current density of 10 mA cm−2 with Tafel slopes of 117, 110, 113, 110 and 113 mV dec−1, respectively. All the doped Ni2P electrodes except Fe-Ni2P decreased the overpotential of the un-doped Ni2P electrode to attain 10 mA cm−2 current density for HER. The 5% Fe-Ni2P electrode with 68 mV overpotential had the best synergistic catalytic performance with the un-doped Ni2P electrode by reducing its overpotential by about 56%. All the doped electrodes had a lower calculated Tafel slope with respect to the un-doped Ni2P electrode suggestive that doping in this study improved the reaction kinetics for HER. Doping Ni2P with Co and Fe only improved slightly the catalytic and reaction kinetics of the Ni2P electrode unlike the OER and supercapacitance. Kucernak et al. fabricated Ni2P and Ni12P5 materials on glassy carbon electrodes and recorded about 270 and 450 mV overpotential with Tafel slopes of 84 and 108 mV dec−1, which are lower than the performance of our electrodes.42 Similarly, Tian et al. prepared Ni12P5 materials from a metal–organic framework that required about 670 mV overpotential to achieve 10 mA cm−2 current density with a Tafel slope of 270 mV dec−1.44 Our as prepared catalysts compare well or outperform other HER catalysts in electrocatalytic examinations and these comparative results are further shown in Table S3.†
The stability and durability of our electrodes for HER were also investigated and are shown in Fig. 11 and S8.† A comparison of the polarization curves for all the electrodes for 1000 cycles are well-matched, indicating excellent durability and stability.
Footnote |
† Electronic supplementary information (ESI) available: TGA of complexes, UV–vis absorption spectra, SEM, TEM, HRTEM and XPS images of nano-materials and comparison tables. See DOI: 10.1039/d1dt01058a |
This journal is © The Royal Society of Chemistry 2021 |