Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Electrocatalytic properties of a carbothermically obtained composite based on vanadium nitridophosphide in the hydrogen evolution, oxygen evolution, and oxygen reduction reactions

Denys O. Mazur, Olena O. Pariiska, Yaroslav I. Kurys* and Vyacheslav G. Koshechko
L.V. Pisarzhevskii Institute of Physical Chemistry of the National Academy of Sciences of Ukraine, 31 Nauky Pr., Kyiv 03028, Ukraine. E-mail: yarkurys@ukr.net

Received 24th July 2025 , Accepted 18th December 2025

First published on 24th December 2025


Abstract

Hydrogen evolution reaction (HER), oxygen evolution reaction (OER), and oxygen reduction reaction (ORR) are the key electrochemical processes of hydrogen energy. Transition metal phosphides, nitrides, carbides, chalcogenides and their composites are promising non-precious metal electrocatalysts for these processes. Catalysts based on so-called early d-metal compounds, particularly V, attract less attention compared to those based on late transition metals (Ni, Co, Fe, etc.), which can generally be explained by their less competitive catalytic characteristics. However, combining V-containing compounds with late transition metal compounds or doping vanadium compounds with late d-metals allows for a significant increase in the performance of V-containing systems. This possibility, as well as the inherent electroconductive properties, ability to function in a wide pH range, and corrosion resistance of a number of vanadium compounds, determine the prospect of their use to create stable and effective electrocatalysts for HER, OER, and ORR. This work firstly shows that pyrolysis of polyaniline doped with phosphoric acid together with sodium metavanadate allows the production composite electrocatalyst for HER, OER and ORR based on vanadium nitridophosphide and N, P-co-doped carbon (V5NP3/N, P-C). The proposed approach is straightforward to implement and avoids the need for ammonia and toxic phosphorus compounds. It was found that in terms of its activity in all the studied processes, V5NP3/N, P-C exceeds the vast majority of known catalysts based on vanadium nitride or phosphide (provided that there are no additional late d-metals in their composition), and is characterized in HER by a Tafel slope (b) of ∼91 mV dec−1 and an overpotential at 10 mA cm−2 (η10) of ∼304 and 325 mV (in 1.0 M NaOH and 0.5 M H2SO4, respectively), in OER – η10 of ∼290 mV and b of ∼241 mV dec−1, in ORR – Eonset of ∼890 mV, E1/2 of ∼820 mV and b of ∼61 mV dec−1. Considering the high functional characteristics of V5NP3/N, P-C in HER, OER and ORR, as a composite based on an early d-metal compound, it can be used, in our opinion, for the formation within the framework of the proposed approach of competitive hybrid electrocatalysts with late d-metals or their compounds.


Introduction

The Hydrogen Evolution Reaction (HER), Oxygen Evolution Reaction (OER), and Oxygen Reduction Reaction (ORR) play a key role in clean and renewable energy technologies.1–3 Specifically, HER and OER provide the basis for electrolytic water splitting – a technology considered the most promising for obtaining high-purity molecular hydrogen due to its environmental friendliness and virtually unlimited water resources. Meanwhile, ORR is implemented in fuel cells and metal–air batteries – devices that provide efficient energy conversion with low environmental pollution. A necessary prerequisite for the practical implementation of HER, OER, and ORR is the use of electrocatalysts to reduce the energy barrier of these reactions.3 Catalysts based on Pt2,4 (HER, ORR) and IrO2 or RuO2 (ref. 5 and 6) (OER) allows for a significant reduction in the overpotential of the respective electrochemical processes, which is caused by electrode polarization. However, despite the efficiency of such catalysts, the high cost of noble metals, their limited resources, and, often, the insufficient stability of the catalysts cause extensive research to find new, much cheaper systems that, in the absence of noble metals, can exhibit high activity in HER, OER, and ORR and stability during long-term operation.

Transition metals (Ni, Co, Fe, Mo, etc.) are much cheaper compared to noble metals and have significant proven reserves. This fact, as well as the high electrocatalytic characteristics in HER, OER, and ORR of a number of d-metal compounds (carbides, nitrides, phosphides, chalcogenides, hydroxides, etc.) established in numerous studies, allows them to be considered as an attractive alternative to noble metal-based catalysts.2,7–16 In this context, the vast majority of research is focused on compounds of so-called late transition metals (particularly Ni, Co, Fe) or their composites. At the same time, catalysts based on early d-metal compounds, particularly V, receive less attention, which can generally be explained by their less competitive catalytic characteristics in HER, OER, and ORR.17 However, combining V-containing compounds with late transition metal compounds (co-catalysts), using the strategy of doping vanadium compounds with late d-metals and non-metal atoms, creating composites with N-doped carbon nanomaterials, etc., allows for a significant increase in the performance of V-containing systems in these processes. This, along with the inherent electroconductive properties, ability to function in a wide pH range, and corrosion resistance (particularly of VN, V carbides, and V phosphides), determines the prospect of their use as stable and effective electrocatalysts for HER, OER, and ORR.17 Considering this, the development of new electrocatalysts based on vanadium compounds is a relevant task.

Recently, we have shown that high-temperature treatment in an inert atmosphere of polyaniline doped with phosphoric acid (PAni·H3PO4) together with a d-metal salt (Co, Ni, Mo, or Fe) allows for the production of composites based on particles of the corresponding d-metal phosphides and carbon co-doped with nitrogen and phosphorus (MetxPy/N, P-C, where Met = Co, Ni, Mo, or Fe) as electrocatalysts for HER, OER, and ORR.18,19 Considering the attractiveness of the proposed approach (simplicity of implementation, low cost of starting reagents, absence of the need to use highly toxic phosphorus compounds), it was of interest to ascertain the possibility of its implementation for obtaining an analogous composite based on vanadium phosphide(s) particles (an early d-metal) and to evaluate the electrocatalytic properties of such a composite in HER, OER, and ORR.

In this work, such a possibility was investigated, and it was shown that the pyrolysis of PAni·H3PO4 (as a source of phosphorus, nitrogen, and carbon) together with sodium metavanadate unexpectedly leads not to the formation of a composite of carbon co-doped with nitrogen and phosphorus with V phosphide(s) particles, but with particles of vanadium nitridophosphide – V5NP3 (V5NP3/N, P-C). In the work, assumptions regarding the formation mechanism of such a composite are made, the main characteristics of V5NP3/N, P-C as an electrocatalyst for HER, OER, and ORR are determined, and a comparison of the electrocatalytic properties of the obtained composite with known catalysts based on vanadium phosphides and nitride is also carried out. To the best of our knowledge, V5NP3-containing composites, as well as the electrocatalytic properties of V5NP3 or hybrid materials based on it, have not been previously discussed in the literature.

Experimental section

Phosphoric acid-doped polyaniline (PAni·H3PO4) was obtained via oxidative polymerization of aniline in an aqueous solution of 2 M H3PO4 (SI).18,20

Synthesis of the composite based on N, P-doped carbon and vanadium nitridophosphide particles. 1.00 g of PAni·H3PO4 was dispersed in 20 mL of ethanol (10 min), and an aqueous solution of 0.28 g NaVO3·H2O (10 wt% metal relative to the doped polymer) was added to the dispersion. The mixture was air-dried, and the dry residue was transferred to a ceramic crucible, which was placed in a tube furnace. The furnace was purged with argon for 15 min, after which the residue was heat-treated in an argon flow at 900 °C for 3 hours, followed by cooling the obtained carbonization product to room temperature in an argon atmosphere. The pyrolysis product was treated with 0.5 M H2SO4 (80 °C, 1 hour), then washed with water on a filter (to pH ∼7) and dried in a drying oven at 90 °C.

N, P-doped carbon (N, P-C) obtained in an analogous manner but in the absence of NaVO3·H2O (SI), as well as Pt/C catalyst (20% Pt on Vulcan XC-72R, FuelCell Store), were used for comparison in electrocatalytic studies.

XRD-patterns of the composite and its precursor were recorded on a D8 ADVANCE diffractometer (Bruker) using filtered CuKα radiation (λ = 0.154 nm). Samples were pressed onto a non-diffracting silicon substrate. Morphology of the composite was investigated by scanning electron microscopy (SEM) using a Tescan Mira 3 LMU microscope (accelerating voltage 10 kV), and its elemental composition was determined by energy-dispersive X-ray spectroscopy (EDX) data (Bruker XFlash detector). The studies of the obtained composites by transmission electron microscopy (TEM) were carried out on a TEM-125K microscope (Selmi), accelerating voltage 100 kV. The Raman spectra of the samples, deposited on a silicon substrate, were recorded on a Renishaw spectrometer (λ = 514 nm). XPS measurements were performed on a PHI 5600 spectrometer equipped with a monochromatic Al Kα source. The instrument work function was calibrated to give the binding energy (BE) of 84 eV for the Au 4f7/2 line for metallic gold and the spectrometer dispersion was adjusted to give a BE of 932.6 eV for the Cu 2p3/2 line of metallic copper. Survey spectra were acquired at a pass energy of 93.9 eV with a step size of 0.2 eV. Prior to analysis, the samples were not pre-etched with the ion gun. Data processing was carried out using CasaXPS 2.3.15 software with the Voigt function (Gaussian/Lorentzian ratio of 70[thin space (1/6-em)]:[thin space (1/6-em)]30). The number of fitted components was limited to the minimum required to describe the spectrum, and their full width at half maximum did not exceed the physically reasonable values for the respective elements. Peak positions were determined with an accuracy of ±0.2 eV. The C 1s photoemission line was used as a reference for energy calibration.

Polarization curves for HER and OER were recorded on a VersaSTAT4-200 potentiostat/galvanostat (Princeton Applied Research) using an undivided electrochemical cell: working electrode – glass carbon (GC) disc with a visible surface area of 0.03 cm2; auxiliary electrode – platinum wire; reference electrode – Ag/AgCl in 3 M NaCl. Polarization curves were iR-corrected. The electrocatalyst loading was ∼1.15 mg cm−2. ORR studies were conducted in the same electrochemical cell on a rotating ring-disk electrode (RRDE 636 A, Princeton Applied Research) in conjunction with a bipotentiostat (MTech BP-10, MTechLab). The visible surface area of the GC disc was 0.2375 cm2, and the Pt ring was 0.1866 cm2. The electrocatalyst loading was ∼0.29 mg cm−2. Current density was determined considering the geometric surface area of the electrode. All potential values are reported relative to the reversible hydrogen electrode (RHE), image file: d5ra05349h-t1.tif.

Preparation of GC electrodes before electrochemical studies involved polishing them with suspensions of diamond particles (1.0 µm) and Al2O3 nanoparticles (0.05 µm), followed by ultrasonic cleaning in a water–ethanol mixture (1[thin space (1/6-em)]:[thin space (1/6-em)]1 vol.) and distilled water. Modification of GC electrodes with electrocatalysts was carried out by applying 2 µL of a dispersion containing 2 mg of the obtained composite (or N, P-C or Pt/C), 48 µL of EtOH, and 8 µL of a 5% Nafion solution in a mixture of low molecular weight aliphatic alcohols, followed by air-drying of the modified electrodes.

Tafel slope (b) was calculated using eqn S1 (SI) from the linear regions of the corresponding Tafel curves. When analyzing data obtained by the RRDE method, eqn S2–S4 (SI) were used to calculate the number of electrons (n) participating in the reduction of 1 oxygen molecule, the hydrogen peroxide yield (H2O2%), and to determine the half-wave potential (E1/2).

The electrochemical impedance spectroscopy (EIS) measurements were performed using a µAUTOLAB III/FRA2 instrument (ECO CHEMIE) in a three-electrode cell. EIS spectra were recorded in the 500 kHz–0.05 Hz frequency range with an amplitude of 10 mV.

Results and discussions

As noted above, the pyrolysis of PAni·H3PO4 with Ni, Co, Mo, or Fe salts yields composites containing phosphide particles of the corresponding d-metals.18,19 However, as shown in the XRD-pattern in Fig. 1a, under identical heat treatment conditions for the precursor based on PAni·H3PO4 and NaVO3·H2O, a composite with a different metal-containing crystalline phase is formed: vanadium nitridophosphide (V5NP3) [PDF 021-0616].21 The obtained composite is hereinafter referred to as V5NP3/N, P-C, considering the obvious presence of carbon in its composition (due to polyaniline carbonization) and the doping of the carbon matrix with nitrogen and phosphorus atoms, which typically occurs during the pyrolysis of N, P-containing polymers in the precursor composition.18,22–24
image file: d5ra05349h-f1.tif
Fig. 1 (a) XRD-patterns of V5NP3/N, P-C and V5NP3 [PDF 021-0616], (b) Raman spectra of N, P-C (1) and V5NP3/N, P-C (2).

Due to the amorphous state of carbon, its presence in the composite is practically undetectable by a characteristic peak in the XRD-pattern at 26.5°. However, it is unequivocally confirmed by the results of Raman spectroscopy on V5NP3/N, P-C. In the Raman spectrum of V5NP3/N, P-C (Fig. 1b, curve 2), similar to N, P-C (Fig. 1b, curve 1), two intense peaks are present with maxima at approximately 1349 and 1596 cm−1. These are characteristics of carbonaceous materials and correspond, respectively, to the disorder-induced phonon mode (D-band), which is caused by defects in the carbon framework, and the graphitic G-band. The relatively high intensity ratio of I(D)/I(G), which is 0.88, indicates a significant degree of disorder (defectiveness) in the carbon within the composite.

Unlike the Raman spectrum of N, P-C, the spectrum of the V5NP3/N, P-C composite shows the appearance of peaks in the mid-frequency region (at 762, 927, and 1003 cm−1) (Fig. 1b). The position and intensity of these peaks suggest they are due to the vibrations of V[double bond, length as m-dash]O, V–O, and/or P–O bonds.25–28 This, in turn, may indicate a partially oxidized state of the V5NP3 particles in the surface layer of the composite. However, according to X-ray diffraction data (Fig. 1a), this oxidation appears to occur only (or predominantly) on the surface of V5NP3/N, P-C and does not extend into the bulk of the composite.

As seen from the presented SEM images (Fig. 2), the obtained V5NP3/N, P-C composite consists of a carbon substrate with a sufficiently developed surface and submicron vanadium-containing particles of irregular shape dispersed on it. These particles primarily form agglomerates approximately 0.5–1.5 µm in size. From the TEM images of V5NP3/N, P-C (Fig. S1, SI) it is seen that the carbon component of the composite is inherently layered, and the majority of the V5NP3 particles (dark contrast) constituting the agglomerates are characterized by sizes of ∼50–100 nm. According to EDX data, the elemental composition (atomic %) of the obtained composite is: V – 1.33; N – 0.90; P – 0.98; C – 86.60; O – 10.19. The atomic ratios of N/V and P/V are greater than what is required to match the established phase composition of the composite, which is consistent with the assumption of N, P-doping of the carbon matrix.


image file: d5ra05349h-f2.tif
Fig. 2 SEM images of the V5NP3/N, P-C composite using backscattered (BSE) (a) and secondary (SE) (b) electrons. Selected area (c) and corresponding elemental mappings for carbon (d), vanadium (e), phosphorus (f) and nitrogen (g).

The SEM selected area and the corresponding EDX elemental mapping images are shown in Fig. 2c–g. The images indicate that the V, P, and N elements are concentrated in regions corresponding to submicron particles distributed on the carbon substrate. This observation is consistent with the XRD data (Fig. 1a) which confirm the presence of vanadium nitridophosphide as a component of the obtained V5NP3/N, P–C composite. Furthermore, the N and P elements are homogeneously distributed over the entire composite surface, suggesting successful doping of the carbon layer with nitrogen and phosphorus.

The presence of oxygen in the elemental composition of V5NP3/N, P-C, consistent with the results of Raman spectroscopy (Fig. 1b), could be a consequence of the composite's surface oxidation due to air exposure.29,30 However, we cannot rule out the presence of a small amount of X-ray amorphous vanadium oxide phases within V5NP3/N, P-C, resulting from the incomplete reduction of oxygen-containing intermediates during the formation of V5NP3 particles via pyrolysis. The possible contribution of oxide impurities to the electrocatalytic properties of V5NP3/N, P-C, when considering the composite as a catalyst for HER, OER, and ORR, was not accounted for by us due to the difficulty of separating the contributions of individual components to the overall catalyst activity.

The surface chemical state of the elements present in the V5NP3/N, P-C was characterized by XPS. The high-resolution spectrum in the O 1s–V 2p region is shown in Fig. 3d. The presence of O 1s core level signal in the V 2p spectra is due to the oxidation of the composite surface, which usually occurs when the sample is in contact with air, including during the XPS test. The peak occurring at binding energy (BE) ∼529.2 eV can be attributed to the O–V bond, and at ∼531.4 eV to the hydroxyl group bonded to VN (O–V/O–H) or to the C[double bond, length as m-dash]O fragments of the oxidized carbon component of the composite.31–33 The V 2p spectrum after deconvolution shows three pairs of peaks (V 2p3/2/V 2p1/2) located at BEs of 514.0/521.6 (V–N/V–P bonding in V5NP3), 515.6/523.2 and 517.5/525.1 eV (Fig. 3b, inset), which are related to V3+ or V4+ (V–N–O/V–P–O), and V5+ (V–O), respectively, and are due to the partial oxidation of the surface of vanadium-containing particles of the composite.31–34 Deconvolution of the N 1s spectrum of the composite (Fig. 3b) results in 3 peaks with maxima at 396.9 (N–V), 399.1 (pyridinic nitrogen) and 401.4 (graphitic nitrogen) eV,34,35 which indicates the participation of nitrogen atoms in the formation of V-containing active sites of the catalyst and doping of the carbon substrate. It should be noted that the peak located in the BE region of ∼399 eV is also often attributed to oxidized nitrogen in the V–N–O fragment.36,37 The P 2p spectrum (Fig. 3c) could be fitted into a doublet at 129.5 and 130.4 eV (corresponding to the binding energy of P 2p1/2 and P 2p3/2, respectively) as well as a peak at 133.2 eV. This doublet can be assigned to P–V bonding, while the high BE peak is attributed to P–O bond in O-containing groups (in particular, PO43−) on the V5NP3 surface.38–40 The high-resolution C 1s XPS spectrum of the catalyst (Fig. 3a) was decomposed into four components that can be attributed to sp2 carbon (∼284.5 eV), structural fragments C–N/C–O/C–P (∼286.7 eV), C[double bond, length as m-dash]O/C[double bond, length as m-dash]N (∼288.9 eV), and also to π → π* satellites of sp2 graphitized carbon (∼291.5 eV).35,41 This state of carbon is explained by the carbonization of the polymer during the formation of the catalyst, doping of the carbon matrix with nitrogen and, possibly, phosphorus atoms, as well as its partial oxidation.


image file: d5ra05349h-f3.tif
Fig. 3 XPS spectra of C 1s (a), N 1s (b), P 2p (c), O 1s and V 2p (inset – the result of the V2p spectrum deconvolution) (d) for V5NP3/N, P-C.

Thus, considering the composition of the precursor and its heat treatment conditions, as well as the phase and elemental composition of the obtained composite, we can propose the following probable formation pathway for V5NP3/N, P-C, which is similar to that previously suggested for composites based on Ni, Co, Fe, and Mo phosphides obtained by an analogous approach.18,19 During the formation of the precursor based on sodium metavanadate and PAni·H3PO4, VOPO4 phosphate is formed, as established by X-ray diffraction method (Fig. S2, SI), due to the interaction between NaVO3·H2O and H3PO4:

3NaVO3·H2O + 4H3PO4 = 3VOPO4 + Na3PO4 + 9H2O

(3NaVO3·H2O + H3PO4 = Na3PO4 + 3H2O + 3HVO3

2HVO3 → V2O5 + H2O

V2O5 + 2H3PO4 = 2VOPO4 + 3H2O)

This is apparently due to the high phosphoric acid content in polyaniline, where it is present not only as a dopant but can also be additionally immobilized within the polymer matrix through specific adsorption.20 With a further increase in pyrolysis temperature above 500 °C, PAni undergoes thermal degradation, leading to the carbonization of the polymer and the release of gaseous products (molecular hydrogen, ammonia).42 Subsequently, via a carbothermal pathway,43,44 the reduction of VOPO4 occurs, resulting in the formation of the “phosphide” component in V5NP3. The product of the reduction of oxygen-containing phosphorus compounds – elemental phosphorus45 – ensures the P-doping of the carbon matrix. Simultaneously, gaseous ammonia, a product of the polymer's thermal degradation, is responsible not only for the N-doping of carbon but also for the formation of the “nitride” component of V5NP3 through ammonification. This latter assumption is supported, in particular, by the known use of other nitrogen-containing polymers – poly-p-phenylenediamine46 and poly-5-aminoindole47 – as the sole nitrogen source in precursors for the formation of the Mo2N phase within N-doped carbon composites under similar heat treatment conditions during their synthesis.

As mentioned above, the electrocatalytic properties of vanadium nitridophosphide or its composites in HER, OER, or ORR had not been previously investigated. Therefore, it was of interest to determine the activity of the obtained V5NP3/N, P-C composite in these processes and compare it with the electrocatalytic characteristics attributed to literature-described catalysts based on phosphides and vanadium nitride (in the absence of additional late d-metals in their composition), as well as N, P-C and Pt/C catalysts.

As a result of studies conducted using linear sweep voltammetry, the ability of V5NP3/N, P-C to exhibit electrocatalytic activity in HER was established in both acidic (Fig. 4a) and alkaline (Fig. 4b) electrolytes. Hydrogen evolution on the V5NP3/N, P-C catalyst occurs at significantly lower potentials than on N, P-C, indicating that the V5NP3 particles are the primary active sites of the obtained composite electrocatalyst in HER.


image file: d5ra05349h-f4.tif
Fig. 4 HER polarization curves and Tafel plots (insets) obtained in 0.5 M H2SO4 (a) and 1.0 M NaOH (b) for GC electrodes modified with V5NP3/N, P-C (1) and N, P-C (2). Potential scan rate: 5 mV s−1.

The onset potential (Eonset) is an important indicator for comparing the efficiency of catalysts in HER. A more effective catalyst for this electrochemical process exhibits a lower (i.e., more positive) Eonset, since less voltage is required for HER initiation. As can be seen from the data presented in Table 1, V5NP3/N, P-C demonstrated a more positive Eonset (was defined as the potential at which the cathodic current density reaches 1 mA cm−2) compared to N, P-C (by 165 mV in 1.0 M NaOH and by 235 mV in 0.5 M H2SO4), indicating its significantly higher activity in HER. For both catalysts, Eonset undergoes a cathodic shift when switching from an acidic to an alkaline electrolyte (Table 1). However, for V5NP3/N, P-C, this shift is much considerably smaller (31 mV) compared to N, P-C (165 mV).

Table 1 Main HER parameters for V5NP3/N, P-C, N, P-C, and Pt/C in electrolytes with different pH
  b, mV dec−1 η10, mV Eonset, mV j0, mA cm−2
0.5 M H2SO4
V5NP3/N, P-C 91 325 −228 2.9 × 10−3
N, P-C 176 640 −463 2.7 × 10−3
Pt/C 21 43 0 1.2 × 10−1
[thin space (1/6-em)]
1.0 M NaOH
V5NP3/N, P-C 91 304 −197 5.9 × 10−3
N, P-C 160 522 −358 5.4 × 10−3
Pt/C 38 21 0 2.3


If we consider the hydrogen evolution overpotential at a current density of 10 mA cm−2 (η10) and the Tafel slope (b) as criteria for catalyst activity, then the efficiency of V5NP3/N, P-C in both aqueous 0.5 M H2SO4 and 1.0 M NaOH is comparable (Fig. 4 and Table 1). In these electrolytes, the catalyst exhibits identical b values and only a minor difference in η10 values (approximately 21 mV). Based on the calculated b value (∼91 mV dec−1), it can be assumed that hydrogen evolution on the V5NP3/N, P-C catalyst proceeds via the Volmer–Heyrovsky mechanism,48 with the Volmer step being the rate-limiting step.

It is worth noting that the difference in η10 and b values between V5NP3/N, P-C and N, P-C is significantly larger in acidic electrolyte (315 mV and 85 mV dec−1) than in alkaline electrolyte (218 mV and 69 mV dec−1) (Table 1). This could be attributed to the higher activity of N, P-C in HER specifically at high pH, primarily due to the presence of CNx-centers in such catalysts.49

Regardless of the electrolyte type, V5NP3/N, P-C is considerably inferior to Pt/C catalyst in terms of both η10 overpotential and Tafel slope. This is not unexpected and is characteristic of catalysts based on vanadium nitride or phosphides when late d-metals are absent from their composition (Table S1, SI). Specifically, the limited productivity of VN in HER is linked to the exceptionally strong Gibbs free energy of hydrogen adsorption (ΔGH*) inherent to this nitride.50,51 It is possible that the same effect occurs for V5NP3/N, P-C. However, the electrocatalytic activity in HER of our obtained composite (considering the calculated η10 and b values) surpasses that of most known analogs based on VN and VP (Table S1, SI). This might be a result of the combination of “nitride” and “phosphide” components within V5NP3/N, P-C, which possibly contributes to optimal hydrogen adsorption on this catalyst, thereby improving its efficiency in HER.

Along with the η10 overpotential and Tafel slope, the exchange current density (j0) is also used to compare the performance of catalysts in HER. As a kinetic parameter, j0 reflects the intrinsic activity of a catalyst under equilibrium conditions. The j0 values for V5NP3/N, P-C and N, P-C catalysts presented in Table 1 were calculated by extrapolating the corresponding Tafel plots to zero overpotential. Since V5NP3/N, P-C exhibits higher j0 values than N, P-C in both electrolytes used, this indicates its higher electrocatalytic activity in HER due to more efficient electron transfer at the electrode/electrolyte interface.

The electrochemically active surface area (ECSA) of V5NP3/N, P-C and N, P-C was estimated based on the double-layer capacitance values of the electrodes modified with the corresponding catalysts calculated from cyclic voltammograms (CVs) recorded at different scan rates (5, 10, 20, 40, 60, 80, and 100 mV s−1) between 0.05 and 0.15 V in 0.5 M H2SO4 and 1.0 M NaOH (Fig. S3a, b, d and e, SI). The potential regions were chosen to avoid interference of pseudocapacitances caused by surface redox reactions. As shown in the Fig. S3c and f a linear correlation is observed when the current density at 0.1 V is plotted against the scan rate. Based on the linear fittings (Fig. S3c, SI) the mass specific capacitance of the catalysts (C) (eqn (1)) was determined, as follows:52

 
C = k/2m (1)
where C is the mass specific capacitance of catalyst, k is the fitting slope (136.8 and 24.3 mF cm−2 in 0.5 M H2SO4, 156.0 and 46.0 mF cm−2 in 1.0 M NaOH for V5NP3/N, P-C and N, P-C, respectively), m is the catalyst areal loading (1.15 mg cm−2).

The C values obtained from eqn (1) are 59.5 (0.5 M H2SO4) and 67.8 (1.0 M NaOH) F g−1 for V5NP3/N, P-C, as well as 10.6 (0.5 M H2SO4) and 20.0 (1.0 M NaOH) F g−1 for N, P-C. Based on these, the ECSA of V5NP3/N, P-C and N, P-C was calculated (eqn (2)).52 The specific capacitance (Cs) (eqn (2)) was chosen as 0.035 (0.5 M H2SO4) and 0.030 (1.0 M NaOH) mF cm−2 based on typical values reported for the indicated electrolytes.53

 
ECSA = C/Cs (2)

Estimated values of ECSA are 170.0 (0.5 M H2SO4) and 226.0 (1.0 M NaOH) m2 g−1 for V5NP3/N, P-C as well as 30.3 (0.5 M H2SO4) and 66.6 (1.0 M NaOH) m2 g−1 for N, P-C. The results show that the ECSA of composite electrocatalyst was much higher than that of N, P-C, indicating that V5NP3/N, P-C had more active sites toward HER, in particular.

In addition, the electrochemical impedance spectroscopy (EIS) measurements were carried out to estimate the charge transfer ability of the obtained catalysts. As shown in Fig. S4a and b (SI) Nyquist plots, the V5NP3/N, P-C possess much lower charge transfer resistance compared to the N, P-C, indicating higher conductivity and faster electron transfer between composite electrode and electrolyte (0.5 M H2SO4 or 1.0 M NaOH). This suggests that the presence of V5NP3 in the composite is beneficial to charge-carrier transportation, thus improving the catalytic performance. The smaller ratio between the sizes of the semicircles for V5NP3/N, P-C and N, P-C in the Nyquist plots recorded in 1.0 M NaOH, compared with those in 0.5 M H2SO4 (Fig. S4a and b, SI), is consistent with the difference in the activity of the obtained catalysts in alkaline and acidic electrolytes (Table 1).

Electrochemical stability is an important criterion for evaluating the performance of electrocatalysts. The long-term durability of V5NP3/N, P-C in HER was estimated by chronopotentiometry test at the constant current density of 10 mA cm−2 in 0.5 M H2SO4 and 1.0 M NaOH. As can be seen from the chronopotentiometry plots (Fig. S5a, SI), overpotential value (η10) for obtained catalyst increases by ∼9% and ∼18% of its initial ones after 8 h of continuous operation in alkaline and acid electrolyte, respectively. On repeated recording of the HER polarization curves after electrolysis for 8 h, a slight cathodic shift is observed for them (Δη10 ∼ 25 and 51 mV in 1.0 M NaOH and 0.5 M H2SO4, respectively) (Fig. S5b and c, SI), indicating minor catalyst degradation.

Alongside HER, the obtained hybrid catalyst V5NP3/N, P-C also exhibits activity in the oxygen evolution reaction (OER) in 1.0 M NaOH (Fig. 5). In this process, it is characterized by Eonset 1.233 V (was defined as the potential at which the anodic current density reaches 1 mA cm−2), η10 value of approximately 290 mV and a Tafel slope (b) of approximately 241 mV dec−1. It's important to note that, according to current understanding, vanadium compounds other than oxides/hydroxides do not directly act as active sites in OER; instead, they serve as precatalysts.17 Under anodic conditions, a shell of amorphous mixed vanadium hydroxides and oxides forms on the V-containing particles (the core),54 and the presence of this shell is what drives the electrocatalytic effect in OER. Considering that the obtained V5NP3/N, P-C provides a significantly lower oxygen evolution overpotential in OER compared to known systems based on VN or VP (Table S2, SI), we can hypothesize several possibilities: a higher efficiency in the formation of the hydroxide/oxide surface layer specifically on the vanadium nitridophosphide particles, or higher electrical conductivity of V5NP3, which accelerates electron transfer between the core and the shell, or a synergistic catalytic effect between the core (V5NP3) and the shell.55 To assess the efficiency of V5NP3/N, P-C and N, P-C in OER, EIS measurements were also performed at a potential of 1.6 V. The resulting Nyquist plots are shown in Fig. S4c (SI). At the measured potential and frequency range N, P-C demonstrates only capacitive behavior, which causes its inefficiency as an OER catalyst. At the same time, a typical semicircle in Nyquist plot due to the charge transfer processes, in particular, occurs for V5NP3/N, P-C. This difference in the shape of the Nyquist plots is consistent with the dramatic differences in the manifestation of activity in OER between V5NP3/N, P-C and N, P-C (Fig. 5a).


image file: d5ra05349h-f5.tif
Fig. 5 (a) OER polarization curves and (b) Tafel plots obtained for GC electrodes modified with V5NP3/N, P-C (1) and N, P-C (2). Electrolyte: 1.0 M NaOH; potential scan rate: 5 mV s−1.

To evaluate the electrocatalytic activity of the obtained hybrid material in ORR, electrochemical studies were conducted using the rotating ring-disk electrode (RRDE) method on GC electrodes modified with V5NP3/N, P-C, as well as, for comparison, N, P-C and Pt/C. The polarization curves for oxygen reduction on these catalysts, obtained in an alkaline electrolyte, are shown in Fig. 6. As the figure illustrates, V5NP3/N, P-C is significantly more effective in ORR compared to N, P-C, indicating the key role of V5NP3 particles in the composite as catalytically active centers. However, as expected, its activity is inferior to that of the Pt-based catalyst.


image file: d5ra05349h-f6.tif
Fig. 6 (a) ORR polarization curves for a GC electrode modified with V5NP3/N, P-C (1) and N, P-C (2) in oxygen-saturated 1.0 M NaOH at a rotation speed of 1200 rpm. (b) Corresponding JrEd dependencies obtained at a Pt-ring polarization potential of 1000 mV; the potential scan rate 5 mV s−1 (c) The H2O2%-Ed and nEd dependences for the V5NP3/N, P-C (1) and N, P-C (2) catalysts.

Table 2 presents the main ORR parameters for V5NP3/N, P-C, N, P-C, and Pt/C catalysts, calculated from RRDE data. Notably, V5NP3/N, P-C as an ORR catalyst exhibits a relatively high Eonset of 0.89 V (was defined as the potential at which the cathodic current density reaches 0.05 mA cm−2) and a half-wave potential (E1/2) of 0.82 V, which are about 50 mV higher than those for N, P-C (Table 2). Another advantage of V5NP3/N, P-C over N, P-C in ORR is a more than twofold lower average hydrogen peroxide yield (H2O2%), which is an undesirable product of this process. Given the established average number of electrons (n) participating in oxygen molecule reduction for the studied catalysts (Table 2), ORR on V5NP3/N, P-C primarily proceeds via a 4-electron mechanism, with a minor contribution from a 2-electron pathway, whereas the contribution of 2- and 4-electron mechanisms to oxygen reduction on N, P-C is comparable.

Table 2 Main ORR parameters for V5NP3/N, P-C, N, P-C, and Pt/C catalysts, calculated from RRDE measurements in oxygen-saturated 1.0 M NaOH at an electrode rotation speed of 1200 rpm
  Eonset, V E1/2, V n H2O2% b, mV dec−1
V5NP3/N, P-C 0.89 0.82 3.58 21.0 61
N, P-C 0.84 0.77 3.09 45.0 38
Pt/C 0.97 0.91 3.96 1.9 51


It is important to note that the hydrogen peroxide yield and, accordingly, n depend significantly on the value of the potential applied to the working electrode (Ed). Moreover, the H2O2%-Ed and nEd dependences for the V5NP3/N, P-C and N, P-C catalysts are fundamentally different (Fig. 6c). At high potentials (∼0.74–0.76 V), the H2O2% and n values are close for both catalysts. However, with an increase in overpotential, there is a sharp increase in the yield of H2O2 in ORR for N, P-C (from ∼32% at E = 0.74 V to ∼57% at E = 0.43 V), while for V5NP3/N, P-C, on the contrary, there is a slight monotonic decrease in H2O2% (from ∼25% at E = 0.74 V to ∼16% at E = 0.43 V). This is in complete agreement with the corresponding nEd dependences (Fig. 6c).

In terms of Eonset and E1/2 values, V5NP3/N, P-C surpasses known ORR catalysts based on VN (Table S3, SI), provided that the latter do not contain additional late d-metals. Compared to VN-containing catalysts, V5NP3/N, P-C also features a lower Tafel slope (b) in ORR (Table S3, SI), which may indicate more favorable kinetics for oxygen reduction, leading to higher productivity in this process. Unfortunately, we haven't found information on the ORR characteristics of systems based on vanadium phosphides, which prevents their comparison with V5NP3/N, P-C as oxygen reduction electrocatalysts.

To characterize the durability of V5NP3/N, P-C in ORR, it was further estimated by chronoamperometry test with the potential holding at 0.6 V in O2-saturated 1.0 M NaOH. As can be seen from the chronoamperometric response (Fig. S6, SI), the V5NP3/N, P-C catalyst retained ∼90% of the initial current after 8 h, thereby showing sufficiently high stability to ORR.

The catalytically active sites involved in the studied processes are fundamentally different in the case of V5NP3/N, P-C and N, P-C. The established significant improvement in the electrocatalytic performance of V5NP3/N, P-C compared to N, P-C may, first of all, be a consequence of the intrinsic activity of vanadium-containing particles that are present in the composite and absent in N, P-C. Also, in the case of OER, the presence of metal-containing particles in the composite determines the possibility of the formation of the hydroxide/oxide layer on their surface, which provides much higher catalytic characteristics in the specified process compared to N, P-C, where this possibility is absent. Although the activity of N, P-C itself in the processes under consideration is low, it is impossible to exclude the positive contribution of the carbon component to the overall activity of the composite, in particular, due to the manifestation of a synergistic effect. The improvement of the electrocatalytic performance of V5NP3/N, P-C compared to N, P-C is also consistent with the established better charge transfer ability and higher ECSA specifically for the composite.

It should be noted that while the electrocatalytic activity of V5NP3/N, P-C in HER, OER, and ORR is moderate, the observed functional advantages of the obtained hybrid material over systems based on vanadium nitride and phosphides as catalysts for these processes (Tables S1–3, SI) suggest that further strategies could significantly enhance its performance. Specifically, incorporating late d-metal doping into V5NP3 within our proposed approach, combining V5NP3 with phosphides (nitrides) of Co, Ni, Fe, and implementing other known modification strategies, could lead to competitive electrocatalytic characteristics for V5NP3-based materials in the discussed reactions.

Conclusions

Thus, we have for the first time synthesized a hybrid composite based on vanadium nitridophosphide and N, P-doped carbon (V5NP3/N, P-C) through the pyrolysis of a precursor made from H3PO4-doped polyaniline and NaVO3. This proposed carbothermal approach is straightforward to implement and avoids the need for toxic phosphorus compounds.

We've established that V5NP3/N, P-C exhibits electrocatalytic properties in HER: (η10 ∼ 304 mV, b ∼ 91 mV dec−1 in 1.0 M NaOH; and η10 ∼ 325 mV, b ∼ 91 mV dec−1 in 0.5 M H2SO4), OER (η10 ∼ 290 mV and b ∼ 241 mV dec−1) and ORR (Eonset ∼ 890 mV, E1/2 ∼ 820 mV, and b ∼ 61 mV dec−1). Furthermore, in the most cases, V5NP3/N, P-C surpasses the activity of known catalysts based on vanadium nitride or phosphides (that lack additional late d-metals) in these processes. This high performance may stem from the combination of “nitride” and “phosphide” components within the obtained composite. It's also been shown that V5NP3 particles play a crucial role as the catalytically active centers in the composite, providing significantly higher efficiency for V5NP3/N, P-C in the discussed reactions compared to N, P-C, which lacks vanadium-containing particles.

Author contributions

D. O. M.: methodology, experimental data production, analysis of the obtained results, and preparation of manuscript materials. O. O. P.: experimental data production and analysis of the obtained results. Y. I. K.: conceptualization, methodology, analysis, and generalization of the obtained results, writing of the original manuscript. V. G. K.: conceptualization, discussion of results, manuscript editing, and guidance.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The data supporting this study's findings are available in the main text and supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05349h.

Acknowledgements

The research was funded by the National Academy of Sciences of Ukraine under project “New nanosized materials and metal complexes, their structure, electrochemical, luminescent and other properties for various functional uses” (no. 0125U000681).

Notes and references

  1. W. Li, H. Tian, L. Ma, Y. Wang, X. Liu and X. Gao, Mater. Adv., 2022, 3, 5598–5644 Search PubMed.
  2. V. Vij, S. Sultan, A. M. Harzandi, A. Meena, J. N. Tiwari, W.-G. Lee, T. Yoon and K. S. Kim, ACS Catal., 2017, 7, 7196–7225 Search PubMed.
  3. T. Ingsel, F. M. de Souza and R. K. Gupta, in Noble Metal-free Electrocatalysts: Fundamentals and Recent Advances in Electrocatalysts for Energy Applications. Volume 1, American Chemical Society, 2022, vol. 1431, pp. 1–29 Search PubMed.
  4. W. Sheng, H. A. Gasteiger and Y. Shao-Horn, J. Electrochem. Soc., 2010, 157, B1529 Search PubMed.
  5. R. Kuriki, O. Ishitani and K. Maeda, ACS Appl. Mater. Interfaces, 2016, 8, 6011–6018 Search PubMed.
  6. V. Pfeifer, T. E. Jones, J. J. V. Vélez, R. Arrigo, S. Piccinin, M. Hävecker, A. Knop-Gericke and R. Schlögl, Chem. Sci., 2017, 8, 2143–2149 Search PubMed.
  7. M. Carmo, D. L. Fritz, J. Mergel and D. Stolten, Int. J. Hydrogen Energy, 2013, 38, 4901–4934 Search PubMed.
  8. X. Zou and Y. Zhang, Chem. Soc. Rev., 2015, 44, 5148–5180 Search PubMed.
  9. M. A. Khan, H. Zhao, W. Zou, Z. Chen, W. Cao, J. Fang, J. Xu, L. Zhang and J. Zhang, Electrochem. Energy Rev., 2018, 1, 483–530 Search PubMed.
  10. C. Hu and L. Dai, Angew. Chem., Int. Ed., 2016, 55, 11736–11758 Search PubMed.
  11. J. Li, Z. Zhao, Y. Ma and Y. Qu, ChemCatChem, 2017, 9, 1554–1568 Search PubMed.
  12. J. Wang, F. Xu, H. Jin, Y. Chen and Y. Wang, Adv. Mater., 2017, 29, 1605838 Search PubMed.
  13. C. Huang, P. Qin, Y. Luo, Q. Ruan, L. Liu, Y. Wu, Q. Li, Y. Xu, R. Liu and P. K. Chu, Mater. Today Energy, 2022, 23, 100911 Search PubMed.
  14. L. Huo, C. Jin, K. Jiang, Q. Bao, Z. Hu and J. Chu, Adv. Energ. Sust. Res., 2022, 3, 2100189 Search PubMed.
  15. N. Mahmood, Y. Yao, J.-W. Zhang, L. Pan, X. Zhang and J.-J. Zou, Adv. Sci., 2018, 5, 1700464 Search PubMed.
  16. W.-F. Chen, J. T. Muckerman and E. Fujita, Chem. Commun., 2013, 49, 8896–8909 Search PubMed.
  17. H. Li, J. Wu, M. Li and Y. Wang, Catalysts, 2024, 14, 368 Search PubMed.
  18. D. O. Mazur, O. O. Pariiska, Y. I. Kurys, V. G. Koshechko and V. D. Pokhodenko, J. Electrochem. Soc., 2024, 171, 076506 Search PubMed.
  19. D. O. Mazur, O. O. Pariiska, Ya. I. Kurys, V. G. Koshechko and V. D. Pokhodenko, Theor. Exp. Chem., 2024, 60, 177–184 Search PubMed.
  20. N. V. Blinova, J. Stejskal, M. Trchová and J. Prokeš, Polymer, 2006, 47, 42–48 Search PubMed.
  21. H. Boller and H. Nowotny, Monatsh. Chem., 1968, 99, 672–675 Search PubMed.
  22. Q. Shi, Q. Liu, Y. Zheng, Y. Dong, L. Wang, H. Liu and W. Yang, Energy Environ. Mater., 2022, 5, 515–523 Search PubMed.
  23. Y. Liu, X. Guan, B. Huang, Q. Wei and Z. Xie, Front. Chem., 2020, 7, 805 Search PubMed.
  24. M. Zhuang, X. Ou, Y. Dou, L. Zhang, Q. Zhang, R. Wu, Y. Ding, M. Shao and Z. Luo, Nano Lett., 2016, 16, 4691–4698 Search PubMed.
  25. P. Shvets, K. Maksimova and A. Goikhman, Phys. B, 2021, 613, 412995 Search PubMed.
  26. T. P. Moser and G. L. Schrader, J. Catal., 1985, 92, 216–231 Search PubMed.
  27. R. L. Frost, K. L. Erickson, M. L. Weier and O. Carmody, Spectrochim. Acta, Part A, 2005, 61, 829–834 Search PubMed.
  28. F. Ben Abdelouahab, R. Olier, N. Guilhaume, F. Lefebvre and J. C. Volta, J. Catal., 1992, 134, 151–167 Search PubMed.
  29. G.-H. An, D.-Y. Lee and H.-J. Ahn, J. Mater. Chem. A, 2017, 5, 19714–19720 Search PubMed.
  30. P. Zhou, D. Xing, Y. Liu, Z. Wang, P. Wang, Z. Zheng, X. Qin, X. Zhang, Y. Dai and B. Huang, J. Mater. Chem. A, 2019, 7, 5513–5521 Search PubMed.
  31. C. M. Ghimbeu, E. Raymundo-Piñero, P. Fioux, F. Béguin and C. Vix-Guterl, J. Mater. Chem., 2011, 21, 13268–13275 Search PubMed.
  32. b Q. Yang, Z. Wang, F. Shi, D. Li, Y. Peng and J. Liu, RSC Adv., 2025, 15, 23994–24001 Search PubMed.
  33. J. Kasten, C.-C. Hsiao, D. Johnson and A. Djire, Nanoscale Adv., 2023, 5, 3485–3493 Search PubMed.
  34. Y. Fu, L. Han, P. Zheng, X. Peng, X. Xian, J. Liu, X. Zeng, P. Dong, J. Xiao and Y. Zhang, Catalysts, 2022, 12, 877 Search PubMed.
  35. N. Zhang, L. Cao, L. Feng, J. Huang, K. Kajiyoshi, C. Li, Q. Liu, D. Yang and J. He, Nanoscale, 2019, 11, 11542–11549 Search PubMed.
  36. H. M. Morales, H. Vieyra, D. A. Sanchez, E. M. Fletes, M. Odlyzko, T. P. Lodge, V. Padilla-Gainza, M. Alcoutlabi and J. G. Parsons, Int. J. Mater. Sci., 2024, 25, 6952 Search PubMed.
  37. M. Chen, H. Fan, Y. Zhang, X. Liang, Q. Chen and X. Xia, Small, 2020, 16, 2003434 Search PubMed.
  38. P. Wei, Q. Geng, A. I. Channa, X. Tong, Y. Luo, S. Lu, G. Chen, S. Gao, Z. Wang and X. Sun, Nano Res., 2020, 13, 2967–2972 Search PubMed.
  39. X. Xu, Y. Lu, J. Shi, X. Hao, Z. Ma, K. Yang, T. Zhang, C. Li, D. Zhang, X. Huang and Y. He, Nat. Commun., 2023, 14, 7708 Search PubMed.
  40. X. Zhang, T. Guo, X. Cao, C. Ma, J. Yan and R. Liu, J. Alloys Compd., 2022, 920, 165932 Search PubMed.
  41. O. Pariiska, D. Mazur, Y. Kurys, R. Socha, V. Koshechko and V. Pokhodenko, J. Solid State Electrochem., 2021, 25, 2309–2319 Search PubMed.
  42. M. K. Traore, W. T. K. Stevenson, B. J. McCormick, R. C. Dorey, S. Wen and D. Meyers, Synth. Met., 1991, 40, 137–153 Search PubMed.
  43. P. Liang, F. Liang, Z. Yao, H. Gao, Y. Sun, B. Jiang and J. Tong, Phosphorus Sulfur Silicon Relat. Elem., 2017, 192, 812–818 Search PubMed.
  44. Z. Yao, H. Hai, Z. Lai, X. Zhang, F. Peng and C. Yan, Top. Catal., 2012, 55, 1040–1045 Search PubMed.
  45. J. Sun, H.-W. Lee, M. Pasta, Y. Sun, W. Liu, Y. Li, H. R. Lee, N. Liu and Y. Cui, Energy Storage Mater., 2016, 4, 130–136 Search PubMed.
  46. Y.-J. Song and Z.-Y. Yuan, Electrochim. Acta, 2017, 246, 536–543 Search PubMed.
  47. Y. I. Kurys, D. O. Mazur, V. G. Koshechko and V. D. Pokhodenko, Electrocatalysis, 2021, 12, 469–477 Search PubMed.
  48. T. Shinagawa, A. T. Garcia-Esparza and K. Takanabe, Sci. Rep., 2015, 5, 13801 Search PubMed.
  49. G. Long, K. Wan, M. Liu, Z. Liang, J. Piao and P. Tsiakaras, J. Catal., 2017, 348, 151–159 Search PubMed.
  50. P. Zhou, D. Xing, Y. Liu, Z. Wang, P. Wang, Z. Zheng, X. Qin, X. Zhang, Y. Dai and B. Huang, J. Mater. Chem. A, 2019, 7, 5513–5521 Search PubMed.
  51. X. Yu, F. Cheng and K. Xie, New J. Chem., 2022, 46, 1392–1398 Search PubMed.
  52. Y. Huang, Q. Gong, X. Song, K. Feng, K. Nie, F. Zhao, Y. Wang, M. Zeng, J. Zhong and Y. Li, ACS Nano, 2016, 10, 11337–11343 Search PubMed.
  53. C. C. L. McCrory, S. Jung, J. C. Peters and T. F. Jaramillo, J. Am. Chem. Soc., 2013, 135, 16977–16987 Search PubMed.
  54. H. Yang, Y. Hu, D. Huang, T. Xiong, M. Li, M.-S. Balogun and Y. Tong, Mater. Today Chem., 2019, 11, 1–7 Search PubMed.
  55. L.-A. Stern, L. Feng, F. Song and X. Hu, Energy Environ. Sci., 2015, 8, 2347–2351 Search PubMed.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.