Open Access Article
Paulin
Kammi Yontchoum
ab,
Carelle Martiale
Kamga Meffo
a,
Bilal
Tasdemir
c,
Kenneth
Mbene
a,
Cedrik
Ngnintedem Yonti
a,
Bilge
Saruhan
c,
Patrice
Kenfack Tsobnang
d and
Roussin
Lontio Fomekong
*a
aDepartment of Chemistry, Higher Teacher Training College, University of Yaoundé I, P.O. Box 47, Yaoundé, Cameroon. E-mail: roussin.lontio@univ-yaounde1.cm
bInstitute of Condensed Matter and Nanoscience, Université catholique de Louvain 1 Place Louis Pasteur, Louvain-la-Neuve 1348, Belgium
cGerman Aerospace Center, Institute for Frontier Materials on Earth and in Space, Linder Hoehe, Cologne 51147, Germany
dDepartment of Chemistry, University of Dschang, P.O. Box 67, Dschang, Cameroon
First published on 24th September 2025
The depletion of fossil fuels and growing environmental concerns have created an urgent global demand for sustainable energy solutions. Among these, hydrogen has emerged as a pivotal energy carrier due to its high energy density and potential for clean combustion. Electrolytic water splitting, especially when powered by renewable energy, offers a promising, carbon-neutral method for hydrogen production. Yet, it is constrained by the slow kinetics of the hydrogen evolution reaction (HER) at the cathode. Thus, obtaining a performant, cheap and environmentally friendly electrocatalyst for HER becomes a big challenge. This study explores the enhancement of HER by developing Cr-doped iron oxide (Fe2O3) nanomaterials, synthesized via a co-precipitation method utilizing succinate ions. Our research demonstrates that Cr-doping significantly modifies the electronic structure and catalytic properties of iron oxide, with optical analyses revealing a bandgap reduction from 1.92 eV to 1.48 eV as Cr3+-concentration increases from 2 to 6.5 atomic percent. The electrocatalytic activity for HER was notably improved, achieving an optimal overpotential of 307 mV at 10 mA cm−2 with a Cr-doping level of 5.5 atomic percent, which outperforms both undoped Fe2O3 and Cr2O3 alone. This performance boost is attributed to the creation of new active sites through lattice defects from chromium substitution into the iron oxide matrix. This investigation not only deepens our understanding of the structure–activity relationships in doped iron oxides but also marks a significant step toward developing efficient, economically viable electrocatalysts for sustainable hydrogen production technologies.
To accelerate the HER, extensive research has been devoted to the development of highly efficient electrocatalysts. Noble metals, such as platinum and iridium, exhibit exceptional catalytic activity for the HER.5 Nevertheless, their scarcity and high cost pose significant challenges for widespread commercialization. As a result, there has been a growing interest in exploring alternative, cost-effective, and abundant materials for electrocatalysis. Transition metal oxides, particularly those based on earth-abundant elements like iron, have emerged as promising candidates due to their tunable electronic properties, high stability, and low cost.
Iron oxides, especially hematite (α-Fe2O3), have been extensively studied as HER and oxygen evolution reaction (OER) electrocatalysts owing to their low cost and abundance.6–8 However, their intrinsic catalytic activity is limited by poor electrical conductivity and a large overpotential. To address these limitations, various strategies have been employed, including doping with foreign atoms to modulate the electronic structure and create additional active sites.9,10 While transition metal-doped iron oxides have been investigated for various applications such as water splitting,11 electrolysis, battery12 and dehydrogenation, the potential of Cr-doped Fe2O3 as an HER electrocatalyst remains largely unexplored. Chromium, with its variable oxidation states and potential for forming stable solid solutions with iron oxide, presents an intriguing candidate for enhancing the HER activity of this material. Previous studies have demonstrated that chromium doping can significantly alter the electronic structure and surface properties of iron oxides, leading to improved catalytic performance in various redox reactions.13,14 Furthermore, the incorporation of chromium may introduce new active sites and modulate the adsorption/desorption kinetics of hydrogen intermediates, thereby facilitating the HER process.
In this study, we report the synthesis and characterization of chromium-doped iron oxide nanomaterials for the alkaline hydrogen evolution reaction. Since the synthesis method is known to have an impact on prepared materials, the co-precipitation method, utilizing succinate ions as a precipitating agent, was employed to synthesize these materials. The selection of the co-precipitation technique is strategic, as it offers precise control over particle size, morphology, and compositional homogeneity, all of which are critical factors influencing the electrocatalytic activity of materials. Moreover, the choice of succinate as a ligand is instrumental in modulating the properties of the resulting oxides. Succinate ions, with their bifunctional character, can coordinate with metal cations to form intermediate complexes, influencing nucleation, growth, and the final crystal structure of the precipitates.15 This approach, compared to other synthesis methods such as sol–gel or hydrothermal, enables the creation of materials with tailored properties, enhancing their potential for the HER. By systematically varying the chromium doping level, we aim to optimize the catalytic activity of these materials for the HER.
To the best of our knowledge, this is the first report on the application of chromium-doped iron oxide synthesized by the succinic acid co-precipitation method as an electrocatalyst for the alkaline HER. This work not only contributes to the fundamental understanding of the structure–activity relationships in these materials but also provides a promising avenue for the development of cost-effective and sustainable hydrogen production technologies.
A series of chromium-doped iron succinate samples were synthesized via co-precipitation in an ethanolic medium. The chromium atomic percentage, relative to the total metal content (Cr/(Cr + Fe)), was systematically varied at 2, 4, 5.5, and 6.5 atomic percent (at%), yielding samples designated as FeCr-Suc1, FeCr-Suc2, FeCr-Suc3, and FeCr-Suc4, respectively. For instance, to prepare FeCr-Suc1, 4.92 mmol of succinic acid ((CH2)2(COOH)2) was initially deprotonated using 9.84 mmol of LiOH.H2O. Subsequently, a mixture of dissolved salts, including 3.21 mmol (0.52 g) of iron(III) chloride (FeCl3) and 0.07 mmol (0.02 g) of CrCl3·6H2O, was added drop by drop in the desired proportions. The mixture was stirred for 90 minutes, filtered, washed, and dried in an oven at 100 °C for 30 minutes, as illustrated in Fig. 1.
Powder x-ray diffraction (XRD) patterns for all synthesized oxide samples were obtained using a Bruker D8 Advance diffractometer (Bruker, Germany) equipped with a Cu Kα radiation source (λ = 1.5406 Å). Data were collected in a 2θ range of 5° to 100° with a step size of 0.05°.
Morphological characterization was carried out using both scanning electron microscopy (SEM) and scanning transmission electron microscopy (STEM). SEM imaging was performed on a Zeiss Ultra 55 microscope (Zeiss, Jena, Germany) operated at an accelerating voltage of 8 kV. Prior to observation, all samples were sputter-coated with a thin platinum layer to minimize charging effects. Complementary high-resolution STEM analysis was conducted using a SM-74240RTED microscope (JEOL, Belgium) operated at an accelerating voltage of 30 kV. Elemental composition analysis was conducted using energy-dispersive X-ray spectroscopy (EDX).
The optical properties of the samples were investigated using a Jasco V750 UV-vis spectrophotometer equipped with an integrating sphere. Absorbance spectra were recorded between 200 and 800 nm in continuous scan mode at a rate of 400 nm min−1.
The measured potentials were converted to the reversible hydrogen electrode (RHE) scale using the Nernst equation:
| ERHE = EAg/AgCl + 0.197 + 0.059 pH |
log|j|, where a is the Tafel intercept, j is the current density, and b is the Tafel slope. Cyclic voltammetry (CV) was performed in the potential range of 0.0 to 0.1 V at various scan rates to determine the electrochemical Cdl. The Cdl was calculated using the following equation: Cdl = (Ja + |Jc|)/2ν, where Ja and Jc are the anodic and cathodic current densities, respectively, and ν is the scan rate. The electrochemical surface area (ECSA) was estimated from the Cdl values using the equation: ECSA = Cdl/Cs, where Cs is the specific capacitance of a planar surface, which was assumed to be 40 μF cm−2, a value commonly adopted in alkaline electrolytes (1 M KOH) as an average reference in the literature.16,17 Electrochemical impedance spectroscopy (EIS) was performed in the frequency range of 0.1 to 100
000 Hz using an AC amplitude of 10 mV. Measurements were recorded at an overpotential corresponding to a current density of 10 mA cm−2. For consistency across all samples, EIS was conducted at −0.307 V vs. RHE, which represents the lowest overpotential required to reach 10 mA cm−2, as determined from the linear LSV curves of the FeCrO-3 catalyst. Finally, chronopotentiometry was employed to assess the long-term stability of the optimal catalyst for 10 hours.
![]() | ||
| Fig. 2 FTIR spectra of pristine Fe-Succ, Cr-Succ and all the Cr-doped Fe succinates (FeCr-Succ1, FeCr-Succ2, FeCr-Succ3 et FeCr-Succ4). | ||
The UV-vis results of all the precursors are as shown in Fig. S1 (see SI). Fe-Succ exhibits a prominent absorption band in the visible region, centered around 500 nm. This is likely due to ligand-to-metal charge transfer (LMCT) transitions involving the succinate ligand and the Fe3+ ion. Cr-Succ shows absorption bands at approximately 420 nm, 579 nm, and possibly a broader one around 709 nm. These bands are likely due to d–d transitions within the Cr3+ ion. The spectra of the mixed succinates (FeCr-Suc1, FeCr-Suc2, FeCr-Suc3, FeCr-Suc4) show absorption features that are a combination of those observed for Fe-Succ and Cr-Succ. They exhibit absorption bands in both the UV and visible regions. Bands at 225 and 270 nm are likely due to π → π* transitions within the succinate ligand. Band at 709 nm could be due to a d–d transition (4A2g → 2T1g) within the Cr3+ ion.19 This absorption behavior reflects the presence of Cr within the Fe-Succ structure. The UV-Vis spectra of the mixed Fe–Cr succinates indicate that the incorporation of Cr3+ into the Fe-Succ structure leads to the formation of new absorption bands in the UV and visible regions. This suggests that the electronic structure of the mixed compounds is different from that of the individual Fe-Succ and Cr-Succ compounds. The presence of Cr3+ ions within the Fe-Succ structure likely perturbs the electronic structure of the Fe3+ ions and the succinate ligands, as observed by the changes in the absorption spectra. The FTIR spectra confirm the bidentate coordination of succinate to both Fe3+ and Cr3+ ions, supporting its role in stabilizing the precursor structure. This coordination could be essential for achieving uniform doping and controlled crystallite formation during thermal decomposition.
![]() | ||
| Fig. 3 (a) FTIR spectra; (b) XRD patterns of pristine Fe2O3, Cr2O3 and all the Cr-doped Fe2O3 (FeCrO-1, FeCrO-2, FeCrO-3 and FeCrO-4). | ||
The FTIR spectra confirm the successful synthesis of Cr2O3 and Cr-doped Fe2O3 nanoparticles. The presence of characteristic vibrational bands indicates the formation of the desired compounds with the expected chemical bonds. The incorporation of Cr into the Fe2O3 lattice is evident from the absence of distinct Cr–O bands, suggesting a well-integrated structure.
The structure state and the crystalline nature of the calcined precursors (Fe-Succ, Cr-Succ, FeCr-Suc1, FeCr-Suc2, FeCr-Suc3, and FeCr-Suc4) were investigated using X-ray diffraction (XRD) analysis, as depicted in Fig. 3b. The XRD patterns of both undoped and Cr-doped samples (Fe2O3, FeCrO-1, FeCrO-2, FeCrO-3, and FeCrO-4) exhibited diffraction peaks corresponding to the (012), (104), (110), (113), (024), (116), and (300) planes of the hematite (α-Fe2O3) phase which peaked at 2θ = 24.2°, 33.1°, 35.6°, 40.9°, 49.5°, 54.1°, and 64°, respectively indexed to the JCPDS standard pattern 01-086-2368 (space group: R
c (167), a = 5.036 Å, b = 5.036 Å, c = 13.75 Å). No additional peaks associated with impurity phases, such as metallic iron or magnetite (Fe3O4), were detected in any sample. Hence the undoped and Cr-doped samples had the same crystal structure and only the α-Fe2O3 phase existed. The average crystalline sizes for all samples were estimated using the Debye–Scherrer formula applied to the (104) peak.
The average crystallite size decreased from 49.14 nm to 38.19 nm (Table 1) with increasing Cr-doping, as evident from the gradual reduction in the intensity of the (104) XRD peak (Fig. S2a). A noticeable shift in the XRD peak position towards higher angles was observed compared to the reference powder (Fig. S2b). The modification observed could be a consequence of the increasing incorporation of smaller chromium(III) ions (ionic radius: 0.615 Å) in place of the larger iron(III) ions (ionic radius: 0.645 Å) as the chromium content increased (Table 1). This substitution process resulted in a contraction of the α-Fe2O3 crystal lattice.21–23 The ensuing lattice shrinkage within the α-Fe2O3 matrix generated increased internal stress and restricted the development of crystallites.24 This result suggests that Cr-doping affects the crystallite structure of α-Fe2O3, as further confirmed by the EDS, electrochemical and optical analyses.
| Sample | 2θ (°) (104) XRD line | FWHM | Crystallite size [nm] |
|---|---|---|---|
| Fe2O3 | 33.06 | 0.17606 | 49.14 |
| FeCrO-1 | 33.22436 | 0.1871 | 46.26 |
| FeCrO-2 | 33.23066 | 0.19341 | 44.75 |
| FeCrO-3 | 33.24588 | 0.22663 | 38.19 |
| FeCrO-4 | 33.24588 | 0.22663 | 38.19 |
| Cr2O3 | 36.2595 (110) | 0.22953 | 38.02 |
Importantly, the XRD patterns of the chromium-doped samples showed no evidence of segregated chromium phases, especially oxides, initially implying that chromium was integrated into the α-Fe2O3 lattice structure.
The morphologies of undoped Fe2O3 and Cr-doped Fe2O3 oxides were examined using SEM and scanning transmission electron microscopy (STEM), with results presented in Fig. 4 and Fig. S3, respectively. These imaging techniques provide complementary insights into the structural evolution induced by Cr incorporation. The SEM images reveal distinct trends in particle size, shape, and agglomeration as Cr content increases.
![]() | ||
| Fig. 4 SEM images of (a) Fe2O3 and Cr-doped Fe2O3, (b) FeCrO-1 (c) FeCrO-2, (d) FeCrO-3 and (e) FeCrO-4. | ||
All samples form dense micro-scale aggregates, likely resulting from thermal-induced agglomeration during the pyrolysis step. The undoped Fe2O3 (Fig. 4a) exhibits a relatively uniform distribution of spherical to slightly elongated particles, with comparatively less agglomeration than the doped samples. As the Cr-doping level increases from FeCrO-1 (Fig. 4b) to FeCrO-4 (Fig. 4e), several morphological changes are observed: a general decrease in average particle size, a transition from spherical to more irregular and angular shapes, and an increase in agglomeration.
These changes suggest that Cr ions disrupt the regular growth of Fe2O3 particles and increase surface energy, promoting aggregation to minimize exposed surface area.
STEM imaging, performed on samples prepared from catalyst ink to reflect operational morphology, confirms and refines these observations. Pristine Fe2O3 (Fig. S3a) forms dense clusters of relatively large particles (80–90 nm), while Cr-doped samples exhibit finer and more dispersed morphologies. FeCrO-1 and FeCrO-2 (Fig. S3b and c) show moderate size reduction and partial clustering, whereas FeCrO-3 (Fig. S3d) presents the most refined structure, with particles averaging ∼ 56 nm and minimal aggregation. FeCrO-4 (Fig. S3e) shows a slight increase in particle size and renewed clustering, suggesting a saturation effect at higher Cr content.
The particle size distribution derived from SEM analysis (Fig. S4) reports average sizes of 82.58 nm (Fe2O3), 76.00 nm (FeCrO-1), 69.51 nm (FeCrO-2), 56.07 nm (FeCrO-3), and 76.80 nm (FeCrO-4). These results corroborate the XRD analysis, which predicts a decrease in crystallite size due to lattice distortion induced by Cr doping. In addition to lattice effects, Cr incorporation likely promotes grain boundary formation, contributing to morphological refinement and stabilization of smaller grains. This hypothesis is supported by atomistic simulations showing that Cr atoms tend to segregate at grain boundaries, reducing their mobility and inhibiting grain growth25 and STEM images (Fig. S3). The resulting defect-rich regions serve as catalytically active sites, enhancing the material's functional performance.26
The elemental mapping analysis presented in Fig. S5 reveals a homogeneous distribution of both Cr and Fe elements within the material. This confirms the successful dispersion of Cr within the Fe2O3 matrix. Energy-dispersive X-ray (EDX) spectroscopy was employed to characterize the elemental composition of all Cr-doped Fe2O3 samples (CrxFe2−xO3). The resulting EDX images, depicted in Fig. S6, unequivocally demonstrate the presence of chromium (Cr), oxygen (O), and iron (Fe) elements. These findings, associated with XRD results, corroborate the successful incorporation of Cr3+ into the Fe2O3 host lattice and the high purity of the synthesized samples. The estimated and are presented in Table 2. A progressive decrease in the atomic percentage of oxygen was observed with increasing Cr-content in the samples, indicating the presence of oxygen vacancies, as confirmed by the FTIR bands and previous work.20 Although pyrolysis was conducted under ambient air conditions, the formation of oxygen vacancies could be explained by several factors. First, the substitution of Fe3+ by Cr3+ introduces local lattice distortion, which lowers the oxygen binding energy and facilitates oxygen release under thermal stress. Second, the elevated temperature during pyrolysis promotes atomic rearrangement and defect mobility, enabling vacancy formation even in an oxidizing atmosphere.27 The observed decrease in oxygen content is directly correlated with a concomitant increase of the percentages of iron within the samples as the Cr-concentration is elevated. Nonetheless, the metal ratios remain remarkably close to the theoretical values.
| Sample | FeCrO-1 | FeCrO-2 | FeCrO-3 | FeCrO-4 | |
|---|---|---|---|---|---|
| Atomic percentage (at%) | O | 57.7 | 52.26 | 48.18 | 45.11 |
| Fe | 41.59 | 46.03 | 49.07 | 51.41 | |
| Cr | 0.71 | 1.71 | 2.75 | 3.48 | |
| Atomic metallic ratio (theorical) | 0.02 | 0.040 | 0.055 | 0.065 | |
| (experimental) | 0.017 | 0.036 | 0.053 | 0.060 | |
![]() | ||
| Fig. 5 (a) UV-visible spectra, and (b) energy bandgap of Fe2O3, Cr2O3 and all the Cr-doped Fe2O3 (FeCrO-1, FeCrO-2, FeCrO-3 and FeCrO-4). | ||
The LSV curves, as illustrated in Fig. 6a, exhibit a reduction in overpotential at a current density of 10 mA cm−2 (η10) as the degree of Cr doping within the Fe2O3 increases.
Fig. 6b reveals that the η10 values for the chromium-doped Fe2O3 samples are notably lower than those for pristine Cr2O3 and Fe2O3, with specific values of 371, 332, 307, and 326 mV recorded for FeCrO-1, FeCrO-2, FeCrO-3, and FeCrO-4, respectively. These are contrasted against a value of 433 mV for undoped Fe2O3 and an almost negligible HER activity for Cr2O3. This suggests that the Cr-doped Fe2O3 samples exhibit superior HER activity, with FeCrO-3 showing the lowest overpotential among the doped variants, likely due to an optimal integration of chromium into the Fe2O3 lattice. Further, SEM and XRD analyses indicated that FeCrO-3 consists of nanoparticles possessing a reduced size in comparison to other doped samples, as detailed in Table 2. This characteristic affords an enlarged surface area for electrocatalytic activities, thereby enhancing the HER performance. That could explain why FeCrO-4 (76.8 nm showed by SEM) has higher potential than FeCrO-3 (56.07 nm showed by SEM)32 Previous scholarly works have documented η10 values of 536 mV for Fe2O3 synthesized via the thermal decomposition of iron(III) malonate.33 The observed differences might be attributed to the choice of ionic ligands used during synthesis. To elucidate the influence of chromium doping on the kinetics of the HER further, Tafel slopes were generated to explore the correlation between overpotential and current density, as depicted in Fig. 7a. The Tafel slope (TS) was determined to be 135 mV dec−1 for Fe2O3, aligning with established literature.33 Conversely, the Cr-doped variants displayed TS values of 84, 118, 130, and 168 mV dec−1 for FeCrO-1, FeCrO-2, FeCrO-4, and FeCrO-3, respectively (Fig. S9).
It is generally desirable for an electrocatalyst to exhibit the lowest possible Tafel slope. The findings suggest that Fe2O3 doped with lower chromium levels (2 at% and 4 at%) significantly reduces the Tafel slope relative to pure Fe2O3. In particular, FeCrO-1, with a Tafel slope of 84 mV dec−1, exhibits improved kinetics for the hydrogen evolution reaction (HER). This implies that minimal Cr doping enhances charge transfer efficiency and boosts HER kinetics. A diminished Tafel slope typically indicates a catalyst's higher efficiency, signifying a quicker reaction rate.
The HER is a two-step electron transfer reaction that occurs at the catalyst surface, proceeding via the Volmer–Heyrovsky or Volmer–Tafel mechanisms. The HER in an alkaline medium involves the following steps:34
Volmer–Heyrovsky mechanism:
| Water dissociation: H2O + * + e− → H* + OH− (Volmer step) |
| Hydrogen desorption: H* + H2O + e− → H2 + OH− (Heyrovsky step) |
Volmer–Tafel mechanism:
| Water dissociation: H2O + * + e− → H* + OH− (Volmer step) |
| Hydrogen recombination: 2H* → H2 (Tafel step) |
According to established HER kinetics, Tafel slopes of approximately 120, 40, and 30 mV dec−1 are typically associated with Volmer, Heyrovsky, and Tafel rate-limiting steps, respectively.35 In this study, the TS values obtained for Fe2O3 and Cr-doped Fe2O3 samples ranged from 84 to 156 mV dec−1. These values initially suggested that the HER mechanism was likely governed by the Volmer step, involving the adsorption of hydrogen intermediates (H*) onto the catalyst surface. This interpretation was subsequently reinforced by the EIS and distribution of relaxation times (DRT) analyses, which revealed slow interfacial charge transfer and adsorption-related processes. Together, these results confirmed that the initial electron transfer remains the rate-limiting step in the HER pathway for these materials.
To delve deeper into the electrode–electrolyte interface kinetics during HER, EIS was performed at the overpotential necessary to achieve a current density of 10 mA cm−2 in 1 M KOH (−0.307 V vs. RHE), as shown in Fig. 7b. The Nyquist plots of all Cr-doped Fe2O3 samples exhibit a single depressed semicircle, characteristic of charge transfer resistance (Rct) associated with HER kinetics.36 Compared to pristine Fe2O3. (Rct ≈ 1300 Ω), Cr-doped samples show significantly reduced Rct values: approximately 280 Ω for FeCrO-1, 790 Ω for FeCrO-2, 990 Ω for FeCrO-4, and notably 50 Ω for FeCrO-3. This confirms that Cr doping enhances interfacial charge transfer, with FeCrO-3 exhibiting the most efficient electron transport among the studied compositions.
The improved EIS performance of FeCrO-3 is attributed to multiple synergistic factors. First, its smaller particle size increases the electrochemically active surface area (ECSA), thereby exposing a greater number of accessible catalytic sites. This morphological refinement facilitates more efficient charge transfer at the electrode/electrolyte interface. Second, Cr-induced lattice distortions and defect formation enhance conductivity by promoting electron mobility across the oxide matrix.
The EIS spectra reveal contributions from solution resistance, charge transfer resistance, and interfacial capacitive effects. While no classical Warburg-type diffusion tail is observed under the present experimental conditions,37,38 a slight inductive response emerges at low frequencies in some samples, particularly FeCrO-1, FeCrO-2 and FeCrO-4. This feature is not a true inductive loop but rather a manifestation of interfacial relaxation phenomena such as intermediate adsorption or surface restructuring of the Fe–Cr oxide layer. Similar low-frequency responses have been reported for HER and OER catalysts in alkaline media.39
To resolve overlapping electrochemical processes that may be obscured in Nyquist plots, a DRT analysis was performed (Fig. S10). By translating the impedance response into a spectrum of relaxation time constants, this approach isolates individual electrochemical processes without relying on any assumed circuit layout.40 In general, all Cr-doped Fe2O3 samples exhibit multiple relaxation features, typically with one peak in the high-frequency region (log
τ ≈ −2 to −3) corresponding to fast charge transfer, and another in the lower-frequency domain (log
τ ≈ 0 to 1) associated with slower interfacial phenomena.40 Among them, FeCrO-3 stands out with the most clearly resolved and intense dual-peak profile. It displays a sharp, narrow peak at log
τ ≈ −2.5, indicative of rapid electron transfer, and a well-defined secondary peak at log
τ ≈ 0.5, reflecting slower surface dynamics. This distinct separation of relaxation times aligns with its exceptionally low Rct and supports its superior HER performance. In contrast, other Cr-doped samples show broader and less resolved DRT profiles, suggesting more complex and less efficient interfacial behavior. The conductivity of the Cr-doped Fe2O3 electrocatalysts was further evaluated through EIS. As Cr content increased, the Rct decreased significantly, indicating enhanced electron mobility within the catalyst matrix. This improvement in conductivity is consistent with previous studies on doped transition metal oxides, where lattice distortion and defect engineering contribute to better charge transport.41,42 These findings reinforce the role of Cr doping not only in modifying the electronic structure but also in facilitating efficient charge transfer during the HER process.
Cyclic voltammetry (CV) was employed to further evaluate the electrocatalytic performance of various materials through the analysis of double-layer capacitance (Cdl) (Fig. S11). This was achieved by conducting CV measurements at varying scan rates ranging from 10 to 100 mV s−1 within the non-faradaic region (0.0–0.1 V). Subsequently, the corresponding electrochemically active surface area (ECSA) was calculated. Fig. 8 illustrates the linear relationship between the current density difference and scan rate at a potential of 0.05 V in a 1 M KOH solution. The Cdl (and the associated ECSA) values for Fe2O3, Cr2O3, FeCrO-1, FeCrO-2, FeCrO-3, and FeCrO-4 were determined to be 444 μF cm−2 (11.1), 52.27 μF cm−2 (1.3), 5140 μF cm−2 (128.5), 3330 μF cm−2 (83.2), 3330 μF cm−2 (83.2), and 1090 μF cm−2 (27.2), respectively.
![]() | ||
| Fig. 8 Scan rate dependence of the current densities at 0.05 V of pristine Fe2O3, Cr2O3 and all the Cr-doped Fe2O3 (FeCrO-1, FeCrO-2, FeCrO-3 and FeCrO-4. | ||
The observed increase in double-layer capacitance (Cdl) from 444 μF cm−2 for pure Fe2O3 to 5140 μF cm−2 for FeCrO-1 upon Cr-doping signifies a substantial enhancement in the number of active sites available for electrochemical reactions. Notably, FeCrO-1 exhibits the highest electrochemically active surface area (ECSA) among all samples. The maximal ECSA and Cdl values observed for FeCrO-1 could likely be attributed to a potentially optimal effect of the low Cr-doping level (2 at%), which might have engendered a significant number of active sites and favorably modified the surface morphology (with lowest particle's agglomeration). This is likely due to the formation of defects within the Fe2O3 lattice when Fe3+ ions are substituted by a smaller Cr3+ ions. These defects (oxygen vacancies) serve as new active sites, thereby increasing the overall surface area available for electrochemical processes.
However, when the amount of Cr-doping exceeds that doping level, as observed in FeCrO-2, FeCrO-3, and FeCrO-4, the particles tend to agglomerate, leading to a reduction in the effective surface area for reactions. This agglomeration is evident from the SEM images and results in a decrease in the number of accessible active sites. Consequently, while there is an initial increase in Cdl with Cr-doping, excessive Cr leads to lower Cdl values due to the reduced availability of active sites resulting from increased particle clustering.
FeCrO-1, which exhibits the lowest Tafel slope (84 mV dec−1) and highest ECSA (128.5), might intuitively be expected to deliver the lowest overpotential. However, this was not observed. A plausible explanation lies in the trade-off between intrinsic site activity, particle morphology, and charge transport dynamics. FeCrO-1, with its lower Cr content and larger particle size (76 nm), possesses highly active sites (Cdl = 5140 μF cm−2) but suffers from limited specific surface area and potentially less efficient charge mobility. In contrast, FeCrO-3, despite having a lower ECSA as indicated by its smaller Cdl (3330 μF cm−2) demonstrates superior catalytic performance. This enhancement is attributed to its smaller particle size (56.07 nm), which increases the number of exposed sites, and more importantly, its significantly lower Rct, reflecting improved electron transfer kinetics. The enhanced conductivity of FeCrO-3 likely stems from Cr-induced lattice distortions and defect formation, which facilitate charge mobility across the electrode–electrolyte interface. Additionally, the spatial distribution of catalyst particles within the Nafion matrix may favor better connectivity and ionic accessibility in FeCrO-3, compensating for its lower ECSA. These synergistic effects collectively contribute to its reduced overpotential and steeper Tafel slope, underscoring the importance of conductivity and interfacial architecture in governing HER performance.
In summary, the data suggest that Cr-doping improves the electrochemical performance of Fe2O3 by increasing the number of active sites up to an optimal level.
The long-term stability of FeCrO-3, which exhibited the lowest overpotential among all Cr-doped Fe2O3 samples, was assessed using chronopotentiometry. The stability test was conducted for 15 hours at a fixed current density of 10 mA cm−2, with the results depicted in Fig. 9a. The overpotential remained essentially constant throughout the test, exhibiting a minimal increase of only 15 mV, indicative of excellent electrochemical stability during sustained operation. This behavior could be attributed to the robust integration of Cr3+ into the Fe2O3 lattice, which enhances defect formation and stabilizes the catalytic surface. The overpotential drift of only 15 mV over 15 h is notably modest. For instance, the electrodeposited CoFe2O4 catalyst reported by Zhang et al. maintained stable activity for several hours of alkaline HER operation, confirming that such low degradation rates are characteristic of robust oxide catalysts.43 To further corroborate the stability of FeCrO-3, a continuous LSV was performed subsequent to the chronopotentiometry test. Post-stability LSV measurements (Fig. 9b) revealed only a minor shift in overpotential at 10 mA cm−2 confirming that the catalyst retained its HER activity without signs of significant degradation. These results demonstrate that FeCrO-3 possesses both high catalytic activity and structural stability, making it a promising candidate for long-term alkaline water electrolysis.
![]() | ||
| Fig. 9 (a) Chronopotentiometry of the best Cr-doped Fe2O3 (FeCrO-3), and (b) its LSV polarization curves before and after chronopotentiometry. | ||
Compared to other nanoelectrocatalysts reported in recent literature,44,45 our Cr-doped Fe2O3 system exhibits competitive overpotential values and stability, while benefiting from a simple and scalable synthesis route.
Table 3 presents a selection of iron oxide-based materials reported as electrocatalysts for the HER in alkaline media. Notably, the FeCrO-3 electrocatalyst synthesized via the straightforward method outlined in this work demonstrates promising characteristics, notably its low overpotential compared to other similar materials and its ability to enhance performance over time. This method, suitable for large-scale production, facilitates molecular-level doping of Fe2O3 with Cr, enhancing the material's dispersibility and catalytic activity.
| Catalyst | η 10 [mV] | Tafel slope [mV dec−1] | Scan rate [mV s−1] | Ref. |
|---|---|---|---|---|
| Niferrite | 600 | 138.9 | 5 | 46 |
| Co0.1Ni0.9ferrite | 566 | 141.1 | 5 | 46 |
| Fe2O3 | 536 | 142 | 5 | 33 |
| ZnFe2O4 | 520 | 144 | 5 | 47 |
| Coferrite | 422 | 116.6 | 5 | 46 |
| NiFe2O4 | 420 | 133 | 5 | 47 |
| Fe 2 O 3 | 433 | 145 | 5 | This work |
| FeCrO-1 | 371 | 128 | ||
| FeCrO-2 | 332 | 143 | ||
| FeCrO-4 | 326 | 144 | ||
| FeCrO-3 | 307 | 164 |
| This journal is © The Royal Society of Chemistry 2025 |