Oxygen vacancy engineering in MXenes for sustainable electrochemical energy conversion and storage applications

Vaishali Sharma a, Jasvir Singh a, Rajnish Dhiman b, Davinder Pal Sharma c and Aman Mahajan *a
aDepartment of Physics, Guru Nanak Dev University, Amritsar 143005, India. E-mail: aman.phy@gndu.ac.in
bDepartment of Physics, Malaviya National Institute of Technology, Jaipur 302017, India
cDepartment of Physics, University of West Indies, St. Augustine, Trinidad and Tobago

Received 13th February 2025 , Accepted 24th June 2025

First published on 10th July 2025


Abstract

Ever-increasing global energy requirements and environmental pollution have directed major research focus on developing sustainable energy conversion and energy storage technologies. Ti3C2TX MXenes are widely considered a potential electrode material for electrochemical hydrogen energy generation through water splitting and electrochemical energy storage supercapacitor applications. Herein, multifunctional Ti3C2TX MXene-based nanocomposites with varying Eu2O3 concentrations were synthesized and systematically investigated for electro/photocatalytic water splitting and two-electrode supercapacitor properties. Ti3C2TX MXenes exhibit high electrical conductivity, tunable surface functionalities and multivalent Ti oxidation states. Meanwhile, the incorporation of Eu2O3 supplements the electrochemical performance by altering the physio-chemical structure and overcoming the intersheet restacking and oxidative degradation issues of pristine Ti3C2TX MXenes. More importantly, the interfacial charge transfer synergism between Ti3C2TX MXenes and Eu2O3 creates oxygen vacancies that modulate the electronic structure of nanocomposites, aiding in the formation of abundant hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) sites, photo-generated charge trapping centres and pseudocapacitive sites. The key findings of the present study showed that the Eu2O3/MXene nanocomposite with an optimum oxygen vacancy content exhibited excellent performance with a small overpotential of 63 mV and 169 mV and a high faradaic efficiency of 96.2% and 95.23% to drive the HER and OER, respectively. Additionally, upon combination with CdS as the photoabsorber, the optimized nanocomposite achieved a high photocurrent density of 4.86 mA cm−2, leading to a H2 evolution rate of 56.67 μL min−1. Considering supercapacitor characteristics, the optimized nanocomposite exhibited a high specific capacitance of 374.98 F g−1 with an energy density and power density of 13.02 Wh kg−1 and 300 W kg−1, respectively. Thus, the results of the present study establish a facile approach to develop high-performance multifunctional electrodes for advancing electrochemical energy conversion and storage technologies.


1. Introduction

Rapid population growth and industrial expansion have placed immense pressure on the overuse of fossil fuels as a primary source of energy production.1 However, their exhausting reserves and subsequent environmental concerns related to the emission of harmful greenhouse gases pose significant threats to climate change and the health of living beings. These adverse effects, coupled with the severe energy shortage crisis, have intensified the exploration of alternative clean and sustainable sources for energy conversion and storage.2

In this context, electrochemical hydrogen (H2) energy generation water splitting devices and electrochemical energy storage supercapacitors stand out as the pioneering pillars of sustainable energy future.3 In the context of electrochemical H2 energy generation, the approach that has gained a rapid upsurge involves the splitting of water into H2 and O2 either via direct electric current (which is termed electrocatalysis) or using sunlight energy (which is termed photocatalysis).4,5 It involves two half-cell reactions, viz. the hydrogen evolution reaction (HER) at the cathode and the oxygen evolution reaction (OER) at the anode, respectively, which control the reaction kinetics and H2 production rate.6 However, practical implementation of this process is greatly hindered by the higher overpotential of existing electrocatalytic materials. Furthermore, limited light absorption, slow charge transport properties and easy electron–hole recombination hamper the photocatalytic water splitting process. In this context, researchers are actively investigating different potential materials and strategies by focusing on these rate-controlling parameters of electro/photocatalytic water splitting. Furthermore, on account of electrochemical energy storage devices, supercapacitors (SCs) have gained significant attention in recent few years due to their intriguing features of quick charge–discharge time profile, high power density, long cyclic stability, safe operation, and compact size.7–9 In SCs, pseudocapacitive materials store nearly 100 times greater charge than electrochemical double layer capacitors (EDLCs) due to the fast redox reaction of the electrolyte with the bulk of the electrode, whereas EDLCs interacts only at the outer interface.10,11 Typically, pseudocapacitive materials alleviate the energy density of SCs due to their wider potential window, but at the cost of sacrificing the power density and cyclability.12,13 However, the widened potential window negatively impacts the overall specific capacitance due to the ongoing HER and OER water-splitting reactions during the charging–discharging processes in SCs.14 Therefore, as the core part of the two electrochemical technologies mentioned above, the material selection is crucial to support the potential window for practicing electro/photocatalytic water splitting and SC applications.

In this direction, researchers have focused on constructing efficient materials with multifunctional catalytic performance.15,16 The conventionally used noble metal-based materials are limited in commercial electrochemical applications due to their high cost and long-term stability issues.17,18 Considering this, MXenes, especially Ti3C2TX MXenes with their notable properties of high electrical conductivity, strong hydrophilic character, high specific surface area, high electrocatalytic activity, and other properties, have become a widely investigated and leading two-dimensional electrode material in the field of electrochemical devices involving catalysis and storage technologies.19,20 It exhibits a rich density of redox sites originating from the transitions of the Ti oxidation states upon reaction with the electrolyte and swift redox activity of the surface-attached functional groups.21,22 Moreover, the high conductivity derived from its carbide core promotes the swift charge transfer kinetics. Nevertheless, the theoretically expected electrochemical parameters, i.e., overpotential of water splitting reactions and specific capacitance of SC have been rarely achieved in experimental analysis mainly due to the oxidative degradation and easy restacking issues of the Ti3C2TX MXene in ambient environment conditions.23,24 The self-restacking of the Ti3C2TX MXene layers due to the strong van der Waals forces reduces the effective surface area of the Ti3C2TX MXene, which suppresses the exposed redox active sites and retards the ion intercalation between the nanosheets.25 This limits the potential window of the pristine Ti3C2TX MXene-based electrodes, and thus degrades their catalytic performance and pseudo capacitance.26–29 Therefore, to derive the high electrochemical performance from the Ti3C2TX MXene, it has been widely proposed to hybridize Ti3C2TX MXene with a promising pseudocapacitive material such as metal oxides, metal hydroxide, and conductive polymers, which modify the physio-chemical structure and induce the efficient interfacial charge transfer properties of Ti3C2TX MXene for higher adsorption characteristics, swift electron transfer and separation properties.30,31

Considering this, metal oxides have attracted much research interest as perspective dopants in Ti3C2TX MXene for their potential use in water splitting and supercapacitors in response to their low cost, easy synthesis routes, environmental friendliness, corrosion resistance, enlarged active surface area and high theoretical capacitance.24,32–34 Aside from the advantages of metal oxides, their intrinsically poor electrical conductivity results in sluggish ion transfer and hampers the power density and cycle life.35–37 In this context, researchers have adopted different strategies to boost the conductivity, such as elemental doping, defect engineering, morphology modification, and heterostructure forming. Among them, the creation of oxygen vacancy (OV) defects is a satisfactory strategy for enhancing the electrical conductivity.38,39 In addition to improving the electrical conductivity, OV offers other advantages in the following ways: (i) OV promotes the adsorption of OH ions, boosting the otherwise sluggish OER and specific capacitance;40,41 (ii) the delocalized electrons of OV modulate the electronic environment of adjacent Ti active sites by reducing the prevalent Ti4+ in Ti3C2TX MXene to Ti2+ for swift HER kinetics; (iii) OV creates additional charge states below the CB, which help to narrow the optical band gap and trap photogenerated electrons, resulting in a significant effect on the photoconversion efficiency;42,43 (iv) the positively charged core of OV surrounded by negatively charged delocalized electrons creates an electric field, which accelerates the interfacial charge transfer characteristics and thus faradaic reactions.44 However, excess OV has a negative impact as recombination centres, hindering electron transfer processes and light absorption. The excess of OV reduces the capacitance by decreasing the available surface area.45,46 Thus, the precise control of OV is crucial for achieving high performance in water splitting and supercapacitor applications.

Towards the controllable incorporation of OV, the investigation of rare-earth metal oxides as an electrode material has gained much attention due to the rapid redox transitions between its mixed-valent oxidation states upon contact with a reducing surface, leading to the formation of OV to maintain charge neutrality.47 Also, these materials exhibit high packing density, which allows for their high volumetric capacitance. Among them, Eu2O3 stands out due to its unique shielded 4f electronic configuration, where the Eu3+/Eu2+ redox couple transitions promote the formation of active reaction sites. Furthermore, the compositing of Eu2O3 with the Ti3C2TX MXene matrix would result in OV formation, which further leads to superior electrochemical performance for electro/photocatalytic water splitting and supercapacitor devices.

Considering the above discussion, in the present work, Ti3C2TX MXene/Eu2O3 nanocomposites were synthesised via a simple sonication method, and the concentration of OV was controlled by varying the weight ratios of Eu2O3 (3%, 5%, and 7 wt%) in Ti3C2TX MXene. The prepared nanocomposites were investigated for HER, OER, overall water splitting H2 energy generation and supercapacitor energy storage device applications. The photocatalytic water splitting properties of these nanocomposites were realized using CdS as the photo absorber material, which demonstrated the efficient impact of the interfacial charge transfer mechanism between MXene and Eu2O3 on the electron/hole pair recombination properties, and thus the photoinduced H2 production efficiency.

2. Experimental

2.1. Materials

Titanium aluminium carbide (Ti3AlC2) MAX powder was purchased from Nanoshel, UK. Europium oxide (Eu2O3) powder, hydrofluoric acid (HF, 40% concentration), sulphuric acid (H2SO4, 98% concentration), Nafion 117 solution, and ethanol were all obtained from Merck, India.

2.2. Synthesis of the MXene/Eu2O3 nanocomposites

Ti3C2TX MXene was prepared using the top-down liquid exfoliation process, as described in our previous reports.48 Eu2O3/Ti3C2TX MXene nanocomposites were prepared using the following procedure: 100 mg of Ti3C2TX MXene powder was dispersed in 20 mL of ethanol and sonicated for 30 min. Afterwards, 3 mg of Eu2O3 powder was added to the above dispersion, and the mixture was stirred for 30 min and later sonicated for another 30 min. After 1 h of reaction, the mixture was left to dry at 50 °C. The dried powder was collected and labelled as 3-EuM. The two other Eu2O3/Ti3C2TX MXene nanocomposites were prepared with 5 mg and 7 mg of Eu2O3, and labelled as 5-EuM and 7-EuM, respectively.

2.3. Characterizations

X-ray diffraction (XRD) measurements were conducted on D8 FOCUS, Bruker Ettlingen with Cu-Kα radiation (λ = 1.54 Å) to collect structural information. The surface morphological visualization of the samples was obtained via field emission scanning electron microscopy (FESEM), Carl Zeiss, Supra 55, and transmission electron microscopy (TEM, JEOL, JEM-2100Plus). X-ray photoelectron spectroscopy data were collected to investigate the chemical composition via Mac-2 electron analyzer using the Mg Kα radiation source.

2.4. Electrocatalytic measurements

The electrocatalytic activity of the prepared samples was analyzed in a standard three-electrode configuration on an AUTOLAB Metrohm PGSTAT302 electrochemical workstation. Pt wire was used as a counter electrode, Ag/AgCl as the reference electrode, and a prepared sample slurry in ethanol deposited on glassy carbon (GC) electrode with an active area of 0.07 cm2 was used as the working electrode. 1 M H2SO4 was used as an electrolyte solution throughout the electrochemical measurements. Cyclic voltammetry (CV) scans were conducted at a scan rate of 10 mV s−1 before performing linear sweep voltammetry (LSV). All measured potentials in this work were converted to reversible hydrogen electrode (RHE), in accordance with the following Nernst eqn (1):
 
ERHE (V) = EAg/AgCl (V) + 0.197 + 0.059pH(1)

LSV measurements were recorded from (0 to −0.4 V vs. RHE) and (1.7 to 1.0 V vs. RHE) for HER and OER, respectively, at a scan rate of 5 mV s−1 to obtain the polarization curves. Electrochemical impedance spectroscopy (EIS) results were obtained in a frequency range of 0.1 Hz to 105 Hz. CV studies were carried out at different scan rates (10 to 80 mV s−1) to obtain data for the double layer capacitance (Cdl) calculation. The electrochemical surface area (ECSA) was derived from Cdl using the following relation (2):49

 
ECSA (cm2) = Cdl/Cs(2)
where Cs refers to the electrode's specific capacitance, which is considered 60 μF cm−2 in the present work.

Furthermore, the roughness factor (RF) was calculated using eqn (3) to estimate the catalytic active sites exposed on the surface.

 
image file: d5ta01195g-t1.tif(3)

The number of active sites (N) was calculated according to the following relation (4):

 
N (mol) = Q/2F(4)
where Q represents the charge in the electrochemical reaction and F refers to Faraday's constant.50

2.5. Photocatalytic measurements

The photocatalytic experiment involved a similar three-electrode configuration in 0.1 M Na2SO3 and 0.1 M Na2S electrolyte. An equal quantity of CdS powder was added to the prepared samples to prepare the photocatalysts. 10 mg of photocatalyst powder was added into a mixture of 15 μL ethanol and 7 μL Nafion solution. The solution was then sonicated for 30 min to form a homogenous slurry, which was drop-casted on a pre-cleaned FTO substrate of 0.5 cm2 active area. The OAI, TriSOL solar simulator with an intensity of 100 mW cm−2 was used as a light source to illuminate the samples and examine their H2 production efficiency under light irradiation. The photocurrent vs. voltage LSV curves were obtained at a scan rate of 5 mV s−1 to determine the photocurrent densities. The photoconversion efficiency (η) was calculated from the LSV curve data using the following relationship (5):
 
η (%) = I(1.23 − ERHE)/I0(5)
where I is the photocurrent density, ERHE is the measured potential, and I0 is the incident light intensity at 100 mW cm−2.51 Mott–Schottky plots were derived at a frequency of 100 Hz to obtain the band edge positions of the photocatalysts.

2.6. Supercapacitor measurements

Electrochemical measurements for supercapacitor device applications were performed using the two-electrode configuration on the AUTOLAB Metrohm PGSTAT302 electrochemical workstation. At first, a thick slurry was prepared by mixing the electroactive material and carbon black in an 85[thin space (1/6-em)]:[thin space (1/6-em)]15 ratio, using absolute ethanol as the solvent, to fabricate the working electrodes. Thereafter, the slurry was doctor bladed onto the aluminum substrates (3 cm × 1.5 cm) and dried for 2 h in an oven. The aluminium substrates were weighed before and after loading the electroactive material to measure the mass loading, which was ∼6 mg. Finally, the two as-prepared identical working electrodes were sandwiched separated with a piece of cellulose filter paper dipped in 1 M H2SO4 (acting as a separator) to obtain the symmetric supercapacitor. Furthermore, the fabricated symmetric supercapacitor was tested using electrochemical techniques, such as cyclic voltammetry (CV), galvanostatic charge–discharge (GCD) and electrochemical impedance spectroscopy (EIS). The specific capacitance of the electrode (Cs,e) and cell (Cs,cell) from CV and GCD curves were evaluated using the following eqn (6)–(9):
 
image file: d5ta01195g-t2.tif(6)
 
image file: d5ta01195g-t3.tif(7)
 
image file: d5ta01195g-t4.tif(8)
 
image file: d5ta01195g-t5.tif(9)

Furthermore, the specific energy (Ecell) and specific power (Pcell) of the cell were evaluated using the calculated value of specific capacitance obtained from the GCD curves, using eqn (10) and (11):

 
image file: d5ta01195g-t6.tif(10)
 
image file: d5ta01195g-t7.tif(11)
where, As is the area under the CV curve, K is the scan-rate (mV s−1), mt is the total mass loading of the cell (g), ΔV is the potential window (V), Im is the current-density (A g−1), and Δt is the discharging time (s).52

The coulombic efficiency (η%) was calculated using the following relation:53

 
image file: d5ta01195g-t8.tif(12)
where, tD is the discharging time and tC is the charging time in seconds.

2.7. DFT studies

First-principles DFT calculations were performed using the Quantum Espresso package.54,55 The Perdew–Burke–Ernzerhof (PBE) approach within the generalized-gradient approximation (GGA) framework was used to describe the exchange–correlation function.56 Spin-restricted plane-wave self-consistent field (PWscf) calculations were performed using ultra-soft pseudopotentials. The cutoff values for the energy and charge density were set equal to 35 Ry and 350 Ry, respectively, and a 10−6 Ry convergence threshold was used.57 To study the MXene system, a slab model of size 3 × 3 × 1 was constructed, which has 63 atoms, and a vacuum of 15 Å was added along the Z direction to avoid the interaction between the adjacent MXene layers.58 Also, to construct the Eu2O3/MXene system, a molecule of Eu2O3 was added above the MXene layer, which is 3 Å away from the layer. The van der Waals force of interaction between MXene and Eu2O3 was taken into account with the help of the Grimme vdW correction (DFT+D3) scheme.59 A Monkhorst–Pack k-point grid of size 3 × 3 × 1 was employed to integrate the Brillouin zone.60 The structural optimizations were carried out by utilizing the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method, and the structure was relaxed to obtain a force (acting on each atom) below the value of 10−3 Ry Bohr−1.61,62 Furthermore, to realize the Eu2O3/MXene system with an oxygen vacancy, one of the oxygen atoms was removed from the MXene layer, and the optimization was then performed.

To study the hydrogen adsorption, an H atom was added to the systems, namely MXene, and Eu2O3/MXene with an oxygen vacancy (Eu2O3/MXene-OV), and then DFT calculations (structural optimizations) were performed. Also, the Grimme vdW correction (DFT+D3) scheme was used to consider the van der Waals interaction between H and the adsorbent system. The hydrogen adsorption energy (Eads) of the adsorbent system has been calculated using the formula:

 
image file: d5ta01195g-t9.tif(13)
where EAdsorbent[thin space (1/6-em)]with[thin space (1/6-em)]H[thin space (1/6-em)]adsorbed is the total energy of the system with hydrogen adsorbed on it, EAdsorbent is the total energy of the system, and EH2 is the total energy of a hydrogen gas molecule.58,63

3. Results and discussion

The X-ray diffraction (XRD) results of Ti3C2TX MXene, 3-EuM, 5-EuM, and 7-EuM are depicted in Fig. 1a. The XRD spectrum of Ti3C2TX MXene shows the typical diffraction peaks located at 2θ = 9.09°, 18.46°, 28.5°, 36.2°, 41.87°, 60.52°, and 72.36°, which are well indexed to the (002), (004), (006), (101), (105), (110), and (311) crystallographic planes of Ti3C2TX MXene, respectively (JCPDS 32-1383).64,65 Furthermore, the diffractograms of all EuM nanocomposites exhibited identical Ti3C2TX MXene diffraction peaks, along with the characteristic XRD peaks of Eu2O3 positioned at 2θ = 18.46°, 28.5°, 32.7°, and 46.5° assigned to the (211), (222), (400), and (440) planes (JCPDS 00-043-1008), respectively.66 Notably, the (211) and (222) XRD planes of Eu2O3 coincide with the (004) and (006) planes of Ti3C2TX MXene. In addition, the intensity of the Ti3C2TX MXene diffraction peaks was found to decrease with a rise in the Eu2O3 doping concentration, which indicates the degeneration of the crystallinity.67 Moreover, it was noticed that the (002) and (004) XRD peaks of Ti3C2TX in the composite samples were shifted to lower 2θ angles at 8.98° and 18.28° in 3-EuM and 8.89° and 18.18° in 5-EuM, suggesting an interplanar lattice expansion of 0.01 nm due to the compositing of Eu2O3 with Ti3C2TX MXene.48 However, in 7-EuM, a slight rise in 2θ was observed for the (002) and (004) XRD peaks to 9.02 and 18.35, respectively, due to the agglomeration effect of Eu2O3 upon excessive doping between the MXene layers.
image file: d5ta01195g-f1.tif
Fig. 1 Structural and morphological analysis of EuM nanocomposites: (a) comparative XRD patterns and (b–d) FESEM images of 3-EuM, 5-EuM, and 7-EuM; (e) TEM image, (f) HRTEM image with inset showing corresponding FFT patterns, and (g) SAED pattern for the 5-EuM sample.

FESEM images were obtained to visualize the morphological features of the EuM nanocomposites, and the results are presented in Fig. 1b–d. The 3-EuM sample (Fig. 1b) exhibits a few Eu2O3 nanoparticles sparsely dispersed on the lamellar Ti3C2TX MXene sheet structure. With an increase in the Eu2O3 doping content in 5-EuM (Fig. 1c), more nanoparticles are seen growing over the Ti3C2TX MXene surface, which effectively prevents the restacking of the Ti3C2TX MXene sheets for the higher electrochemically accessible specific surface area and electron diffusion efficiency. Importantly, the Ti3C2TX MXene sheets prevented the aggregation of the Eu2O3 nanoparticles due to their layered structure, thus maximizing the nanoparticle potential merits. However, with the further increase in the doping content of Eu2O3 in 7-EuM (Fig. 1d), the lamellar structure of Ti3C2TX MXene gets distorted, leading to the formation of a thicker MXene sheet structure with the collapse of the effective pore channels, causing the severe blockage of the potential redox active sites. It reduces the specific surface area, which is considered to negatively impact the electrochemical application prospects.

Furthermore, to examine the elemental composition and uniform distribution of elements throughout the materials, energy-dispersive X-ray spectroscopy (EDS) measurements were conducted and elemental mapping was obtained. The EDS spectra (Fig. S1a–c) indicate the presence of Ti, O, C, F, and Eu elements in the EuM nanocomposites, where the content of the Eu element was found to increase in the 3-EuM, 5-EuM and 7-EuM samples due to the increasing doping concentration of Eu2O3. This is evident from their respective elemental distribution table, which demonstrates the wt% and atomic% concentration of each element in the samples. Furthermore, the EDS elemental mapping shows the distribution pattern of elements in the nanocomposites, which reveal the uniform dispersion of the Eu2O3 nanoparticles on the MXene surface, thus confirming the formation of EuM nanocomposites. A more detailed microstructural observation of optimized 5-EuM was noted using TEM investigations. Consistent with the FESEM results, the TEM image (Fig. 1e) shows Eu2O3 nanoparticle growth on the Ti3C2TX MXene sheets morphology. The corresponding high-resolution TEM (HRTEM) micrograph (Fig. 1f) demonstrates the construction of numerous heterointerfaces between Ti3C2TX MXene and Eu2O3, which depicts the distinctly visible parallel ridges. Fast Fourier Transform (FFT) studies carried out on the ridges revealed an interlayer spacing of 0.31 nm and 0.28 nm (insets: Fig. 1f), matching well with the (222) and (101) planes of Eu2O3 and Ti3C2TX MXene, respectively.68 The slightly enlarged lattice spacing of Ti3C2TX MXene in the nanocomposite, relative to the pristine Ti3C2TX MXene of 0.27 nm, is attributed to the increased intersheet distance upon insertion of the Eu2O3 nanoparticles.69 Furthermore, the selected area diffraction (SAED) pattern of the 5-EuM nanocomposite (Fig. 1g) represents a combination of bright diffraction spots and circular rings assigned to the single crystalline Ti3C2TX MXene structure and polycrystalline Eu2O3 nanoparticles, respectively.48,70

XPS was carried out to determine the oxidation states of the nanocomposites upon interfacial interactions between the doped Eu2O3 and Ti3C2TX MXene surface. The XPS survey spectra (Fig. S2) of the EuM composites depict the presence of Ti, O, C, F, and Eu with increasing intensity of Eu 4d from the 3-EuM sample to the 7-EuM sample, confirming the formation of the composite with no detectable impurity peaks. The Ti 2p core-level XPS spectra (Fig. 2a–c) display two spin–orbital doublets, viz., Ti 2p3/2 and Ti 2p1/2 at 459.1 eV and 464.9 eV, respectively. Further deconvolution of the Ti 2p spectra provides detailed oxidation states, i.e., Ti–C (454.8 eV), Ti2+ (456.5 eV), Ti3+ (458.5 eV), and Ti4+ (459.1 eV).71 The atomic (at) ratios of these species in the nanocomposites have been calculated from their relative peak area and are represented in Table 1. It is seen that with an increase in concentration from 3-EuM to 5-EuM, Ti2+/Ti4+ increases from 0.9 to 4.9, suggesting that the effective reduction of the Ti species of Ti3C2TX MXene is related to the OV creation. Specifically, the formation of OV creates two free electrons, which reduces the Ti4+ states to Ti2+ states.72 However, with the further increase in concentration in 7-EuM, Ti2+/Ti4+ exhibits a decrease, which could be attributed to the agglomeration impact of the excessive concentration, resulting in the interfacial charge transfer bottleneck.73 In addition, a similar trend was noticed upon analyzing the detailed Eu 4d spectra (Fig. 2d–f), in which peaks at B.E. of 128.82 eV and 137.21 eV could be assigned to Eu2+ and Eu3+.74 A comparison of the Eu 4d spectra revealed that the Eu2+/Eu3+ content increased from 0.08 in the 3-EuM sample to 0.14 in the 5-EuM sample, indicating that a portion of Eu3+ was transformed into Eu2+ states, and consequently created OV to maintain the charge neutrality. Furthermore, the high-resolution O 1s XPS spectra (Fig. 2g–i) were probed to estimate the OV contents. It depicts three distinct peaks positioned at 530.3 eV, 531.8 eV, and 533.2 eV associated with lattice oxygen (OL), surface oxygen vacancies (OV), and surface adsorbed water H2Oads, respectively.75 As tabulated in Table 1, the OV concentration increased to 2.61 in the 5-EuM sample, which is indicative of the increased OV content, while a sharp rise was seen in the 7-EuM sample, which is ascribed to the excessive generation of oxygen vacancies upon higher doping levels.


image file: d5ta01195g-f2.tif
Fig. 2 High-resolution XPS spectra of (a–c) Ti 2p, (d–f) Eu 4d, and (g–i) O 1s for the 3-EuM, 5-EuM, and 7-EuM nanocomposites.
Table 1 Atomic percentages (at%) of the valence states for the 3-EuM, 5-EuM, and 7-EuM nanocomposites
Sample Ti2+/Ti4+ Eu2+/Eu3+ OV (%) H2Oads
3-EuM 0.91 0.08 27.90 25.83
5-EuM 4.91 0.14 41.42 42.69
7-EuM 1.8 0.11 42.89 30.60


Thus, it is inferred that the precursor concentration greatly influences the structural, morphological, and composition of the synthesized samples. The structural and morphological analyses definitively confirmed the anchoring of nanoparticles on the Ti3C2TX MXene sheets during ultrasonic treatment. The XPS results confirmed the Eu2O3 loading and revealed the at% of the Ti, Eu species and oxygen moieties. The overall analyses show that among the prepared samples, 5-EuM exhibits greater potential in electrochemical energy conversion and storage applications, owing to its structural suitability, larger hydrophilicity, large density of Ti2+ and Eu2+ as the potential HER active sites, optimal OV content which would ease otherwise sluggish OER pathway, and multivalent oxidation states with greater OH ion adsorption ability favoring charge storage ability.

Considering the above discussion, the electrocatalytic HER performance of EuM nanocomposites was first investigated from the cathodic polarization curves (Fig. 3a) obtained by conducting linear sweep voltammetry (LSV) measurements in 1 M H2SO4 electrolyte. For comparison, a state-of-the-art HER catalyst, Pt/C, was also tested among the prepared catalysts. In the HER LSV polarization study, Pt/C clearly exhibited excellent performance in HER activity, requiring a minimal overpotential (η) of 47 mV to attain 10 mA cm−2. Among the prepared synthesized materials, it has been observed that the 5-EuM sample exhibits near Pt activity, presenting a lower overpotential of 63 mV, and was found to be catalytically more active than the 3-EuM and 7-EuM samples (Table 2). The excessive generation of OV in 7-EuM negatively impacts the electroactivity by reducing the electrical conductivity and imposing structural instability.76 This is because the higher OV content creates a larger defect that localizes a high electron density around the defect site, thereby hindering the material's charge transfer ability and increasing the overpotentials.77 Removing the excessive lattice oxygen atoms distorts the crystal lattice, which affects the active site and adsorption/desorption process. More importantly, the highly defective structure is prone to corrosion under electrochemical testing conditions, thereby accelerating the degradation process.75 Therefore, the performance of the 7-EuM sample with the highest oxygen vacancy concentration was lower as compared to that for 5-EuM. The superior electrochemical performance of 5-EuM suggests that the controlled generation of OV played a vital role in enhancing the HER catalytic activity. To further obtain insight into the reaction kinetics and understand the underlying reaction mechanism on the catalyst surface, the Tafel plot was extracted from the corresponding LSV curves. The reaction catalyzed by Pt/C, 3-EuM, 5-EuM, and 7-EuM showed Tafel slopes (Fig. 3b) of 45.0 mV dec−1, 134.1 mV dec−1, 65.4 mV dec−1, and 125.1 mV dec−1, respectively. Aside from the superior catalytic activity, the stability of an electrocatalyst in its electrolyte environment is also crucial to ensure its long-term operation in water electrolysis for commercial-scale H2 production. Thus, to probe the stability of the 5-EuM electrocatalyst, chronoamperometric (CP) electrochemical testing (Fig. 3c) was conducted at an overpotential of 63 mV for a continuous 50 h. It revealed that 5-EuM retained a nearly steady current, revealing its outstanding durability. A slight catalytic decrement with time can be ascribed to the bubble accumulation and release from the electrode surface, which obstructs the effective electrolyte interaction with the electrode surface. This electrochemical stability is further confirmed from the negligible change in the overpotential observed in the LSV measurement performed before and after conducting 100 CV cycles (inset: Fig. 3c), signifying no obvious catalyst degradation.


image file: d5ta01195g-f3.tif
Fig. 3 Comparison of the electrochemical performance of the as-synthesized EuM nanocomposites for the HER and OER in 1.0 M H2SO4: (a) LSV polarization curves, (b) corresponding Tafel slopes for Pt/C, 3-EuM, 5-EuM, and 7-EuM; (c) chronoamperometry (it curve) stability test for 5-EuM for 50 h; (d) LSV polarization curves and (e) corresponding Tafel slopes of RuO2, 3-EuM, 5-EuM, and 7-EuM; (f) chronoamperometry (it curve) stability test for 5-EuM for 50 h.
Table 2 Electrochemical HER and OER parameters for the 3-EuM, 5-EuM, and 7-EuM nanocomposites
Electrocatalyst HER OER
η (mV) Tafel slope (mV dec−1) η (mV) Tafel slope (mV dec−1)
3-EuM 205 134.1 237 184.0
5-EuM 63 65.4 169 92.0
7-EuM 155 125.1 173 140.0


It is well known that electrocatalysis is highly pH-dependent.78 Thus, to compare the electrocatalytic water splitting kinetics of the 5-EuM sample in acidic, alkaline, and neutral solutions, LSV measurements were conducted in 1 M H2SO4 acidic, 1 M KOH alkaline, and 1 M PBS neutral media (Fig. S3). In comparison to acidic electrocatalysis, relatively deteriorated HER performance is seen in alkaline and neutral solutions with higher overpotentials of 81.2 mV and 141.9 mV, respectively, indicating higher HER kinetics in acidic media. This is primarily attributed to the higher H+ ion concentration in acidic media (H2SO4 and H2O), which are involved in the electro-reduction processes towards H2 generation, whereas the limited H+ availability in alkaline and neutral electrolytes (derived from H2O only) restricts the electrocatalysis kinetics.79

The OER activity of the prepared samples was also evaluated in 1 M H2SO4. RuO2, a benchmarking OER catalyst, was also investigated for its water oxidation ability. The OER LSV polarization curves in Fig. 3d show that 5-EuM requires an η of 169 mV, which is even lower than that required for the RuO2 catalyst (η = 220 mV), demonstrating the superior activity of the EuM composites, especially 5-EuM in the oxidation reaction. The derived Tafel slopes (Fig. 3e) for RuO2, 3-EuM, 5-EuM, and 7-EuM are found to be 184.0 mV dec−1, 92.0 mV dec−1, and 140.0 mV dec−1, respectively. The smaller η and Tafel slope values of 5-EuM as the OER catalyst are in accordance with the HER results, which signify the strong redox properties of the 5-EuM sample towards both H2 and O2 evolutions. This remarkable performance can be ascribed to the presence of the optimal OV content in the 5-EuM sample calculated from the XPS results, which significantly promotes the rate of charge transfer even at lower applied η values. Likewise, CP tests conducted at 169 mV OER overpotential (Fig. 3f) and long-term cycling stability tests (inset: Fig. 3f) further demonstrated the stable OER kinetics over a long duration of 50 h.

In addition to the long-term stability measurements, post-catalytic structural and morphological characterizations were obtained for the optimized 5-EuM sample to validate the electrode durability. The post-stability PXRD pattern of the 5-EuM sample (Fig. S4a) indicated no significant structural alterations, demonstrating a diffraction pattern that was similar to that obtained before electrochemical testing. Likewise, the SEM image (Fig. S4b) indicated the dispersion of the Eu2O3 nanoparticles while preserving the layered morphology of MXene, further revealing the high stability of the sample after the continuous stability test. This could be attributed to the high corrosion resistance of the dispersed Eu2O3 rare-earth oxide, which encapsulated the MXene surface in aqueous solution and protected it from oxidative degradation during acidic electrolysis.

Electrochemical double layer capacitance (Cdl) and electrochemical surface area (ECSA) were obtained to further analyze the intrinsic electrochemical behaviour of the electrocatalysts. ECSA is directly related to Cdl, which is measured by conducting CV in a non-faradaic potential range at different scan rates from 10 to 80 mV s−1, as shown in Fig. 4a–c. The linear fitting of the scan rate vs. current density plot (Fig. 4d) gives the Cdl values (Table 3). The Cdl of the 5-EuM sample (11.35 mF) is much larger than that of the other tested electrodes, indicating the presence of a rich density of active sites at the electrode/electrolyte interface, which is responsible for ion adsorption and interaction. Consequently, the ECSA calculated from the corresponding Cdl values gives the largest electrochemically accessible surface area for the 5-EuM sample (189.1 cm2), validating a greater number of adsorbed species on the catalyst surface. The number of active sites (N) were calculated from the Cdl values to estimate the density of the surface exposed redox active sites. Clearly, 5-EuM gives the highest concentration of catalytically active sites (11.7 × 10−8 mol), evidencing its high HER performance. The roughness factor (RF) was evaluated further to compare the HER performance of the prepared sample. The 5-EuM sample revealed the highest RF (2701.4), again confirming the presence of abundant actives sites, which is beneficial for the higher H2 evolution.


image file: d5ta01195g-f4.tif
Fig. 4 (a–c) CV curves at different scan rates (10–80 mV s−1), (d) Cdl plots, and (e) Nyquist plots fitted using the equivalent circuit for the 3-EuM, 5-EuM, and 7-EuM nanocomposites. (f) LSV curve of the overall electrochemical water splitting for 5-EuM before and after the cycling test for a two-electrode setup.
Table 3 C dl, ECSA, RF, N and Rct values for 3-EuM, 5-EuM, and 7-EuM nanocomposites
Electrocatalyst C dl (mF) ECSA (cm2) RF N (10−8 mol) R ct (Ω)
3-EuM 3.78 63.0 900 3.91 45.91
5-EuM 11.35 189.1 2701.4 11.7 7.13
7-EuM 7.66 127.6 1822.85 7.93 27.81


The conductivity of an electrocatalyst is an important parameter to evaluate the electroactivity. Thus, EIS measurements were carried out to determine the resistance offered to the flow of charges during an electrochemical reaction. The Nyquist plots (Fig. 4e) obtained after fitting the EIS data to an equivalent circuit (inset: Fig. 4e) present a semicircle with a diameter corresponding to the internal charge transfer resistance (Rct), revealing that 5-EuM delivers an obvious smallest Rct of 7.13 Ω in comparison to the 3-EuM and 7-EuM samples. This signifies the improved scope of the electrode/electrolyte interfacial contact for 5-EuM, owing to the synergistic effects of the electron-rich OV, Eu2O3 nanospacers, and highly conductive Ti3C2TX MXene support, which are responsible for its enhanced performance.

Finally, 5-EuM was assembled as an anode and cathode in an overall water-splitting electrolyzer setup. As shown in the polarization curves (Fig. 4f), the 5-EuM catalyst delivered a current density of 10 mA cm−2 at a small cell voltage of 1.48 V, revealing the remarkable bifunctionality of the prepared catalyst. This could be ascribed to the presence of OV in a controlled manner, which not only modulates the electronic environment of the neighbouring Ti valence states of Ti3C2TX MXene for generating highly active Ti2+ HER sites, but also that they act as adoption centers for OER intermediates and enhance the charge transfer kinetics. The water-splitting polarization curves before and after the repeated CV scans have been recorded (Fig. 4f). The results indicate the outstanding cycling stability, as observed from the negligible cell voltage drift recorded after continuous 100 CV cycles.

Finally, the faradaic efficiencies for HER and OER of 5-EuM were evaluated using the experimentally quantified volumes of H2 and O2 obtained at the cathode and anode, respectively, in a typical water-gas displacement process conducted at a constant overpotential in chronoamperometric measurement and theoretically calculated volumes. After 3600 s, a total of 8.96 C of the charge passing through the electrolyzer cell generated 2 mL and 0.99 mL of H2 and O2, which agrees well with the theoretically estimated volume of 2.079 μL and 1.040 μL for H2 and O2, respectively. This gives the faradaic efficiency of 96.2% and 95.23% for HER and OER, respectively.

Furthermore, the photocatalytic water-splitting performance of the synthesized composites was investigated by combining the nanocomposites with a potential photoabsorber, CdS. LSV was first carried out under artificial sunlight illumination. Then, the current density vs. voltage (IV) curves were obtained to compare the current density change with the voltage upon light illumination. Fig. 5a demonstrates that the photocurrent density is altered significantly upon varying the doping concentration. The maximum photocurrent density for 3-EuM/CdS, 5-EuM/CdS, and 7-EuM/CdS at 1.0 V vs. RHE reached 0.3 mA cm−2, 4.86 mA cm−2, and 2.07 mA cm−2, respectively. The remarkable increment of the photo response in 5-EuM/CdS is attributable to the promoted charge separation efficiency, which generated a greater number of e/h+ pairs with high charge transfer ability. Beyond this doping concentration in 7-EuM/CdS, the decrease in the current density indicates the diminished charge carrier density, which could be directly correlated with a greater e/h+ pair annihilation rate over charge separation. This is because the higher OV concentration in 7-EuM/CdS acts as photogenerated charge recombination centers and reduces the charge mobility, which negatively effects the photocatalytic performance. Furthermore, the photocurrent efficiency was calculated from the corresponding photocurrent spectra to quantitatively estimate the photocatalytic H2 generation efficiency of the nanocomposites. Fig. 5b shows that 5-EuM/CdS exhibits η% of 0.034%, which is higher than that for 3-EuM/CdS (0.002%) and 7-EuM/CdS (0.02%). Thus, consistent with the LSV results, the η% analysis also reveals that the 5-EuM sample should be considered optimal for achieving high photoelectrochemical activity, which is due to the desired physio/chemical structure modulation induced by OV formation. The higher concentration of Eu2O3 reduced the photoactivity, which could be related to the creation of excessive OV that act as charge recombination centers. Thus, it is concluded that when the OV concentration is too low, it becomes impossible for electrons to reach the electrode surface and participate in surface chemical reactions, whereas an excessive OV concentration increases the charge recombination and thus degrades the photocatalytic activity. Therefore, the optimal OV concentration is crucial for promoting the photoexcited charge carrier separation and transfer properties.


image file: d5ta01195g-f5.tif
Fig. 5 (a) LSV polarization curves, (b) photoconversion efficiency, and (c) Nyquist plots. (d) UV-vis absorption spectra, (e) Tauc plots, and (f) PL spectra for the 3-EuM/CdS, 5-EuM/CdS, and 7-EuM/CdS photocatalysts.

Additionally, EIS was conducted to analyze the charge transfer kinetics at the photoanode/electrolyte interface and the corresponding Nyquist plots are shown in Fig. 5c. Among the prepared photocatalysts, 5-EuM/CdS exhibited the smallest semi-circular arc radius of 16.37 Ω when compared to 3-EuM/CdS (41.89 Ω) and 7-EuM/CdS (21.87 Ω). This signifies that the incorporation of Eu2O3 in the controlled content and conductive Ti3C2TX MXene support greatly diminished the resistance towards the charge transport.

Importantly, the use of Pt wire as the counter electrode is very challenging, especially for HER activity since Pt itself is a catalyst. Therefore, to explore the efficiency of the prepared catalyst towards electrochemical and photoelectrochemical testing, the measurements were collected using a graphite rod as the counter electrode under similar experimental conditions.

The HER polarization curves for electrocatalytic water splitting (Fig. S5a) and photocatalytic water splitting (Fig. S5b) demonstrated no significant shift in the overpotential in electrocatalytic water splitting. A nearly similar current density in photocatalytic water splitting was attained while using a graphite rod compared to the Pt counter electrode, indicating that the HER contribution is mainly derived from the intrinsic catalytic activity of the prepared Eu2O3/MXene sample. This could be attributed to the high conductivity of the 5-EuM sample derived from the electron-rich oxygen vacancies, compensating for the otherwise deteriorated conductivity upon using a graphite rod instead of the Pt counter electrode.

To explain the significantly enhanced photocatalytic properties of the 5-EuM/CdS sample, the light absorption behaviour and charge separation efficiency of all samples were comparatively investigated using UV-vis reflectance spectroscopy and PL studies, respectively. Fig. 5d presents the UV-vis absorption spectra of the 3-EuM/CdS, 5-EuM/CdS, and 7-EuM/CdS photocatalysts. Notably, the absorption band edge of 5-EuM/CdS is greatly shifted to the higher wavelength region, which suggests the enlarged light absorption capability. The increase of the visible range absorbance of 5-EuM/CdS is ascribed to the formation of the intra-band states of the OV defects below the conduction band, which alter the band structure to expand the light absorption response. Also, the presence of delocalised free electrons in OV creates electron charge clouds, which absorb incident light energy and undergo intra-band excitation processes. Nevertheless, the excess OV concentration disrupts the band structure, where the excessive electrons accumulated in the forbidden band hinder the light absorption. Considering this, the band gaps of the nanocomposite samples were estimated from the Tauc plots (Fig. 5e) obtained by using Kubelka–Munk theory on the corresponding DRS spectra. Upon extrapolation of the linear portion of Tauc plot to the horizontal axis, the band gap values are noted as follows, i.e., 2.30 eV, 2.05 eV, and 2.10 eV for 3-EuM/CdS, 5-EuM/CdS and 7-EuM/CdS, respectively. Furthermore, the PL spectra of the doped samples were obtained at 520 nm to study the recombination rate of the photo-generated e/h+ pairs and to better understand the role of the underlying OV formation on the photocatalytic properties of the samples. As illustrated in Fig. 5f, all of the samples exhibited an emission peak around 687 nm, and the PL intensity was found to decrease from the 3-EuM sample to the 5-EuM sample, revealing the greatly diminished charge recombination. This indicates the positive aspect of OV in the effective charge separation and transfer characteristics of the conductive Ti3C2TX MXene support, which elongates the charge carrier lifetime. With the further rise in the Eu2O3 concentration in the 7-EuM sample, the PL intensity greatly increased due to the swift recombination rate of the e/h+ pairs, showing the negative impact of excessive OV, which behave as recombination centers. The combined UV and PL analysis confirms that the creation of OV in a controlled concentration does not significantly increase the light harvesting efficiency, but suppresses the charge carrier recombination by trapping photogenerated e/h+ pairs, improving the interfacial charge transfer properties, and thus enhancing the photocatalytic efficiency.

Furthermore, the photocatalytic H2 evolution plots (Fig. 6) were obtained to confirm the superiority of 5-EuM/CdS. Fig. 6a and b shows the maximal photocatalytic efficiency of 3.4 mL in 60 min over the 5-EuM/CdS photocatalyst, which shows a linear trend over the time course. As a result, 5-EuM/CdS exhibited the highest average H2 evolution rate of 56.67 μL min−1, which presents remarkably improved photocatalytic performance compared to 3-EuM/CdS (38.34 μL min−1) and 7-EuM/CdS (45 μL min−1) shown in Fig. 6c. Thus, it is evidenced that the improvement in the photocatalytic water splitting for 5-EuM/CdS primarily originates from the optimal OV concentration by the controlled Eu2O3 doping content, which efficiently alleviated the charge recombination processes. In addition, the photocatalytic durability of 5-EuM/CdS was assessed under similar reaction conditions to explore the viability for long-term photocatalysis application. The results from Fig. 6d showed that 5-EuM/CdS, when recycled three times over 60 min, exhibited a nearly stable H2 generation rate, indicating its high operational stability.


image file: d5ta01195g-f6.tif
Fig. 6 (a) Photocatalytic H2 evolution curves over a time course, (b) comparison of the photocatalytic activity, and (c) average H2 evolution per min for 3-EuM/CdS, 5-EuM/CdS, and 7-EuM/CdS. (d) Stability test for 5-EuM/CdS cycled three times.

Based on the above discussed results, the plausible charge transfer and redox reaction mechanism for the enhanced photocatalytic performance can be proposed. Firstly, the band gap energies (Eg) for CdS and the Eu2O3 components are estimated to be 2.36 eV and 4.23 eV, respectively, using UV-vis (Fig. S6a and b) and the corresponding Tauc plots (inset: Fig. S6a and b). Furthermore, Mott–Schottky plots were obtained to determine the energy band structure of the semiconductor photocatalyst. The positive slope of the Mott–Schottky plots (Fig. S6c and d) confirms that both CdS and Eu2O3 are n-type semiconductors. The extrapolation of the slope of Mott–Schottky plots to the x-axis gives a flat band potential (Vfb) of −0.29 and −0.19 V vs. RHE for CdS and Eu2O3, respectively. ERHE can be further converted into a vacuum energy level (Evac) using eqn (14):

 
Evac = −4.5 − ERHE(14)

Generally, in n-type semiconductors, the conduction band (CB) lies at 0.2 eV above Vfb. Meanwhile, in p-type semiconductors, the valence band (VB) is positioned at 0.2 eV below the Vfb. Thus, considering the band gap values, the CB/VB positions for CdS and Eu2O3 are −4.01/−6.42 eV and −4.11/−8.41 eV, respectively. Considering these derived band edge positions of CdS and Eu2O3, the energy level structure with Ti3C2TX MXene as a co-catalyst is shown in the schematic (Fig. 7). Upon light irradiation, both CdS and Eu2O3 components are simultaneously excited by absorbing light of energy greater than their respective band gap values. In CdS, the electrons are excited from their valence bands and migrate quickly to fill the CB, leaving behind holes in its VB and creating an electron/hole pair. Meanwhile, as confirmed from XPS analysis, the presence of OV in Eu2O3 forms a defect energy level just below its CB minima edge, lowering the band gap for broader light absorption capability.80 Moreover, these OV act as trapping centers for photoexcited electrons from CB and prevent them from recombining with holes in VB. This process greatly minimizes the electron/hole recombination and facilitates their effective separation in bulk Eu2O3.42 Furthermore, the close intimate contact between the two semiconductors causes the electrons in the CB of Eu2O3 to recombine with the holes present in the VB of CdS via a Z-scheme charge transfer route. Consequently, the photogenerated electrons in CB of CdS and photogenerated holes in VB of Eu2O3 are spatially separated on their respective band energy levels with strong reduction and oxidation ability, respectively. Typically, the local electric field generated across the positive core and negatively charged delocalized electron cloud of OV accelerates the swift charge transfer for effective interfacial charge transfer. Furthermore, when CdS and Ti3C2TX MXenes are in contact with each other, the difference in their Ef position causes the electrons to flow from the CB of CdS to Ti3C2TX MXenes until an equilibrium state of Ef of both CdS and Ti3C2TX MXenes is reached. This raises the Ef of Ti3C2TX MXene to a suitable reduction potential, i.e., more negative than 0 V vs. NHE. After Ef alignment, a space charge layer is created on the CdS side, which causes the upward band bending of the CdS energy levels. This leads to the creation of a Schottky barrier at the metal–semiconductor interface region, which prevents the reflux of electrons, thus promoting the unidirectional electron flow to MXene. Thus, Ti3C2TX MXene now acts as an electron sink, where these electrons effectively reduce the Ti valence states for higher H+ ion adsorption and release of molecular H2. Meanwhile, holes accumulated on VB of Eu2O3 exhibits high oxidisability due to its VB position at the more positive potential than the standard water oxidation potential, i.e., 1.23 V vs. NHE. Thus, the holes effectively oxidise the water molecules to generate O2. It is revealed that interactions among CdS, Eu2O3, and Ti3C2TX MXene could effectively promote the interfacial charge separation and boost the charge transfer efficiency of the photogenerated electron/hole pairs.


image file: d5ta01195g-f7.tif
Fig. 7 Schematic illustrating the photocatalytic water splitting mechanism.

To study the energy storage supercapacitor device applications, electrochemical analysis including CV, GCD, and EIS measurements were carried out in a two-electrode sandwiched cell configuration with 1.0 M H2SO4 aqueous electrolyte. Initially, the CV response of Ti3C2TX MXene and EuM nanocomposites were compared at 10 mV s−1 (Fig. 8a). It was found that all of the nanocomposites achieved a wider operating voltage window (0 V to 1 V) in comparison to Ti3C2TX MXene (0 V to 0.6 V) under the same scan rate, which holds a direct correlation with the energy density. For Ti3C2TX MXene, the CV curve depicts neither an exact rectangular profile (EDLC characteristic) nor distinct redox peaks. The plausible reason for this behaviour may be ascribed to the interlayer restacking issues of Ti3C2TX MXene, which first inhibit the intercalation/deintercalation movement of the electrolyte ions, obstructing the charge storage via EDLC mechanism. Secondly, the interlayer collapse reduces the specific surface area and thus masks the redox active sites, inhibiting the charge storage through reversible surface redox reactions. The slight humps are only observed due to the interaction of hydronium ions with the –O functional entity on the Ti3C2TX MXene surface, which alters the Ti oxidation state in an acidic medium and thus imparts a slight pseudo influence. Thus, the shape of the CV curve of Ti3C2TX MXene suggests the quasi-pseudo behaviour. Meanwhile, on compositing with Eu2O3, the CV curves of the EuM nanocomposites display a relatively larger CV peak area and prominent pairs of redox peaks, which reflects the higher specific capacitance (Cs) compared to that for the pure Ti3C2TX MXene, evidencing the pseudo capacitance behaviour. The EDLC contribution in capacitance could also be seen from the nearly flat middle section of the CV curve. Thus, the combination of pseudo-capacitance and EDLC behaviour contributes to the nanocomposites' overall electrochemical energy storage capacity. This could be ascribed to the in situ generation of OV upon incorporation of Eu2O3 in Ti3C2TX MXene. The positive core of OV acts as an active site that favors adsorption and accumulation of OH on the electrode surface. It modulates the local electronic structure of the underlying Ti3C2TX MXene Ti sites towards the active redox ability for pseudo-capacitance. Simultaneously, OV boosts the electrochemical charge transfer and expansion of the interlayer diffusion channels of Ti3C2TX MXene with Eu2O3. Considering the vital role of OV, the 5-EuM sample exhibits the highest current density and largest CV integrated area, signifying the relatively higher charge storage capacity. Moreover, the 5-EuM CV curve exhibits more prominent redox peaks relative to other nanocomposites, which is directly correlated with multiple redox reactions attributable to the greater number of active sites on the electrode surface, along with swift redox kinetics due to the optimal OV content and reversible transitions between Eu2+/Eu3+. The smaller OV level offers less redox active sites; meanwhile, an excessive OV concentration imposes a negative impact by reducing the conductivity.81,82 Thus, the Cs calculated from the CV plots using eqn (6) and (7) were 135.46 F g−1, 201.95 F g−1, 364.35 F g−1, and 245.98 F g−1 for Ti3C2TX MXene, 3-EuM, 5-EuM, and 7-EuM, respectively.


image file: d5ta01195g-f8.tif
Fig. 8 (a) CV curves and (b) GCD curves comparison for Ti3C2TX MXene, 3-EuM, 5-EuM, and 7-EuM.

Furthermore, GCD measurements were employed to confirm the enhanced electrochemical performance of the 5-EuM sample. Fig. 8b shows the symmetric GCD profiles of the prepared electrodes, which were collected at 0.6 A g−1 and compared. 5-EuM revealed greatly improved capacitance with the Cs reaching as high as 374.98 F g−1, higher than Ti3C2TX MXene (139.88 F g−1), 3-EuM (206.28 F g−1), and 7-EuM (246.98 F g−1). Moreover, a comparison of the GCD curves demonstrates a much longer discharging time of 156 s for 5-EuM. These results indicate that a suitable amount of Eu2O3 increases the charge storage ability of Ti3C2TX MXene by contributing to the pseudo capacitance and increasing the interlayer spacing between the Ti3C2TX MXene layers.

Furthermore, the CVs and GCDs of 3-EuM, 5-EuM and 7-EuM were recorded at different scan rates (10 to 80 mV s−1) and current densities (0.6 to 1.4 A g−1), as shown in Fig. 9. It can be seen from the CV studies that with increasing scan rate, the current response increases, while Cs decreases (Table 4). With the increasing scan rate, the electrolyte ions do not have enough time to interact with the electrode's surface. Thus, the specific capacitance value is reduced at higher scan rates. The shape of the GCD studies at higher current density represents an exact triangular shape. Meanwhile, as the current density decreases, GCD demonstrates a slow and sloppy discharge, elongating the discharge time and thus resulting in an increase in Cs (Table 5). This may be due to the deep penetration of ions at the slower current density, which resulted in faradaic reactions at the inner surface and additional charge storage. Slower discharge arises from the formation of soluble redox shuttles upon decomposition of the loosely bonded surface functional groups at lower current density. The highest current response, largest specific capacitance, and longest discharging time altogether for the 5-EuM sample confirm its suitability for charge storage in supercapacitor applications.


image file: d5ta01195g-f9.tif
Fig. 9 (a, c, and e) CV curves at different scan rates and (b, d, and f) GCD curves at various current densities for the 3-EuM, 5-EuM, and 7-EuM samples.
Table 4 Comparison of the cell and electrode specific capacitance values at different scan rates for 3-EuM, 5-EuM, and 7-EuM obtained from CV studies
Scan rate (mV s−1) Cell specific capacitance from CV studies (F g−1) Electrode specific capacitance from CV studies (F g−1)
3-EuM 5-EuM 7-EuM 3-EuM 5-EuM 7-EuM
10 50.49 91.09 61.49 201.95 364.35 245.98
20 34.41 75.26 55.24 137.65 301.05 220.95
40 22.51 51.99 38.69 90.04 207.98 154.78
60 20.16 43.63 32.95 80.42 174.52 131.82
80 17.99 36.67 29.05 71.99 146.68 116.18


Table 5 Comparison of the cell and electrode specific capacitance values at different scan rates for 3-EuM, 5-EuM, and 7-EuM obtained from GCD studies
Current density (A g−1) Cell specific capacitance from GCD studies (F g−1) Electrode specific capacitance from GCD studies (F g−1)
3-EuM 5-EuM 7-EuM 3-EuM 5-EuM 7-EuM
0.6 51.57 93.74 61.75 206.28 374.98 246.98
0.8 37.01 75.14 59.13 148.03 300.54 236.54
1.0 26.51 53.32 41.61 106.04 213.28 166.44
1.2 22.63 46.03 36.85 90.53 184.13 147.41
1.4 20.41 36.69 30.40 81.65 146.78 121.58


Furthermore, the energy densities and power densities were calculated from the GCD parameters of 3-EuM, 5-EuM, and 7-EuM, and compared in Ragone plots (Fig. 10a). At every power density value, 5-EuM depicts the highest energy density of 13.02 Wh kg−1 at 0.6 A g−1 current density among the other tested samples, which is attributable to the increased density of exposed redox active species upon creation of OV at a controlled concentration. Another observation suggests that the energy density is increased upon doping without sacrificing the power density due to the increased electrical conductivity and ion diffusion rate associated with the delocalized free electrons of OV, leading to high power density. Furthermore, EIS revealed that 5-EuM exhibited superior interfacial charge transfer and capacitive behaviour compared to the bare Ti3C2TX MXene and other nanocomposites. The Nyquist plots (Fig. 10b) derived from EIS measurements could be divided into three major components: the internal resistance (Rs) of the electrode material, electrolyte ionic movement and metal electrode are obtained from the intersection of the plot with the X-axis in the high-frequency region; the diameter of the semicircular part gives the charge transfer resistance (Rct) at the electrode/electrolyte interface; and the linear part in the low-frequency region represents the Warburg resistance in which the slope of line corresponds to the diffusion/transportation of electrolytic ions. Table 6 lists the Rs and Rct values calculated for different electrodes. The larger the Rs and Rct, the greater the charge transfer kinetics. Furthermore, the steep line indicates the facile electrolyte intercalation, and thus higher capacitive performance. Among the prepared samples, 5-EuM displayed the smallest Rs and Rct of 1.58 Ω and 3.86 Ω, respectively, and a nearly vertical line in the low-frequency region, which projects its fast kinetic merits. Finally, a long-term cycling stability test (Fig. 10c) was also conducted for 5-EuM by repeated GCD tests at 1.0 A g−1 current density. It maintained 93.39% of specific capacitance even after 10[thin space (1/6-em)]000 continuous cycles, which demonstrates the durability of the electrode material. The coulombic efficiency of the 5-EuM sample at a current density of 1 A g−1 has been calculated from the GCD curves in order to understand the material's performance during the long-term charge–discharge cycles. The coulombic efficiency of the 5-EuM sample decayed from 91.69% at the 1st cycle to 90.31% at the 10[thin space (1/6-em)]000th cycle after continuous measurement (inset: Fig. 10c). This suggests the high stability of the MXene electrode materials, which is attributed to the high corrosion resistance properties upon composite formation of Eu2O3.


image file: d5ta01195g-f10.tif
Fig. 10 (a) Ragone plot, (b) Nyquist plot; inset shows the magnified high-frequency region for the 3-EuM, 5-EuM, and 7-EuM electrodes, and (c) cycling stability of 5-EuM at 1.0 A g−1.
Table 6 EIS parameters for Ti3C2TX MXene, 3-EuM, 5-EuM, and 7-EuM
Samples R s (Ω) R ct (Ω)
Ti3C2TX MXene 3.604 13.34
3-EuM 2.07 5.97
5-EuM 1.58 3.86
7-EuM 1.68 4.24


A comparison of the water splitting and supercapacitor performance of Ti3C2TX MXene/Eu2O3 in this present work and the previously reported MXene-based electrochemical devices is tabulated in Table S1. It is easy to see that the present device exhibits lower overpotential values in HER and OER and higher specific capacitance due to the optimal generation of OV, which aids in swift charge transfer, modulates an electronic structure that creates abundant redox active sites, maximizes the light absorption, and favors greater OH ion adsorption for higher charge storage capacity.

The electronic structure of a material plays a vital role in the HER process. Thus, DFT calculations were carried out to calculate the electronic structure and hydrogen (H) adsorption energy of MXene and the Eu2O3/MXene heterostructure with an oxygen vacancy. The optimized structure of the MXene system is shown in Fig. 11a. The optimized structure of MXene with an H adsorbed on the O-site is shown in Fig. 11b, and the calculated value of H adsorption energy is equal to −0.48 eV. A high value of H adsorption energy (−0.48 eV) on the O-site suggests a strong interaction (coupling) between the H and O atoms and subsequent strong adsorption. However, this strong adsorption tendency reflects the difficulty in the desorption of H from the surface of the adsorbent (MXene), consequently hindering the HER process. These findings are also supported by the DFT studies from other researchers.58,63,83,84


image file: d5ta01195g-f11.tif
Fig. 11 (a) Top and front views of the optimised structure of MXenes. (b) Top and front views of H adsorbed on the O site of MXenes. (c) TDOS and PDOS plots of MXenes.

Furthermore, OV are formed in the Eu2O3/MXene system, and play a vital role in the electrocatalytic process.63 Therefore, the adsorption/desorption of hydrogen and the electronic structure of Eu2O3/MXene-OV were also investigated. The structure of the Eu2O3/MXene-OV system before and after optimization is shown in Fig. 12a. To realize the Eu2O3/MXene-OV system, OV has been created by removing an O atom from MXene. However, after structural optimization, the OV shifted to the Eu2O3 group, as shown in Fig. 12a. The H adsorption on the Eu2O3/MXene-OV system (Fig. 12b) is calculated to −0.084 eV, which is favorable for both the adsorption and desorption of H, and thus for the HER process.58,63,84 Hence, DFT studies confirm that the presence of OV in the Eu2O3/MXene system helps to improve the HER performance.63


image file: d5ta01195g-f12.tif
Fig. 12 (a) Top and front views of the structure of the Eu2O3/MXene system with an OV before and after optimisation. (b) Top and front views of H adsorbed on the O site of the Eu2O3 molecule of the Eu2O3/MXene-OV system. (c) TDOS and PDOS plots of Eu2O3/MXene-OV.

Furthermore, the electronic structure (density of states) helps to study the charge carrier transfer rate.85 The total density of states (TDOS) plot of MXene and the partial density of states (PDOS) plot of the constituent atoms (Ti, C, O) are shown in Fig. 11c. The top of the valence band is formed from the hybridization between the orbitals of C, Ti, and O. In contrast, the bottom of the conduction band is mainly made up of Ti orbitals, with some contribution from the orbitals of C and O.86 It should be noted that while plotting the DOS plots and band structures, the Fermi energy level was set at 0 eV. The DOS and PDOS plots of Eu2O3/MXene-OV are shown in Fig. 12c. Apart from the contribution of Ti, C, and O of MXene, the orbitals of Eu mainly contribute to the conduction band, whereas O (of the Eu2O3 group) primarily contributes to the valence states particularly close to the Fermi level. From the TDOS plots, we can see that there are a greater number of electronic states around the Fermi level in the case of Eu2O3/MXene-OV as compared to Eu2O3/MXene, which indicates that the electron transfer rate is higher in Eu2O3/MXene-OV. In other words, the larger the number (density) of electronic states near the Fermi level, the higher the concentration of electrons around the Fermi level, which leads to better electron mobility and hence higher electron transfer rate.85 For a better comparison of the density of states of the MXene and Eu2O3/MXene-OV near the Fermi level, see Fig. S7.

Thus, the DFT study reveals that Eu2O3/MXene-OV demonstrates better adsorption/desorption energetics, greater electron transfer rate, and reaction kinetics, making it a suitable contender for electrochemical energy conversion and electrochemical energy storage applications.

4. Conclusion

In summary, Eu2O3/Ti3C2TX MXene electrodes with tunable oxygen vacancies (OV) were prepared and tested for electro/photocatalytic water splitting and supercapacitor device applications. FESEM, TEM and XPS studies revealed the physio/chemical alterations in the Ti3C2TX MXene structure with OV variations upon changing the concentration of Eu2O3. OV modulated the electronic structure of Ti3C2TX MXene by providing highly active Ti2+ sites for effective H+ reduction, and eased the sluggish charge dissociation process in OER. Among the prepared samples, the 5 wt% Eu2O3-doped Ti3C2TX MXene (5-EuM) with optimal OV content demonstrated superior performance in electrocatalytic water splitting, requiring smaller 63 mV and 169 mV overpotential to drive HER and OER, respectively, and also exhibited long-term stable operation in an acidic medium for 15 h. Furthermore, 5-EuM assembled in a two-electrode configuration required a small cell voltage of 1.48 V to achieve 10 mA cm−2 current density for the overall water splitting reaction. Upon combination with the CdS photoabsorber, the 5-EuM/CdS sample exhibited superior performance in photocatalysis with a higher photocurrent density of 4.86 mA cm−2 among the other photocatalysts, and a high photoconversion efficiency of 1.86%. This enhanced photocatalytic response could be ascribed to the maximized light absorption and an effective charge separation mediated by the controlled OV content in 5-EuM/CdS and suitable redox potentials of the adjoining photocatalysts, leading to an efficient Z-scheme interfacial charge transfer. When used as a supercapacitor electrode, 5-EuM delivered a high specific capacitance of 374.98 F g−1 and a long-discharging time of 156 s derived from the combined effect of the redox-active Eu2O3 and highly conductive underlying Ti3C2TX MXene matrix. Thus, the present work paves the way for tuning the electrochemical catalytic activity and charge storage by utilizing the synergistic effects of the OV-enriched Eu2O3/Ti3C2TX MXene framework and constructed multifunctional electrodes, exhibiting superior electrochemical energy generation water splitting and electrochemical energy storage supercapacitor applications.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

Vaishali Sharma: conceptualization (equal); data curation (lead); methodology (lead); writing – original draft (lead). Jasvir Singh: investigation (equal); resources (equal); Rajnish Dhiman: investigation (equal); Aman Mahajan: supervision (lead); conceptualization (equal); writing – review & editing (equal).

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors are thankful to UGC-DAE CSR, Indore, for providing financial support and experimental facility through CRS project no. CRS/2021–22/01/386.

References

  1. P. Wang, S. Zhang, Z. Wang, Y. Mo, X. Luo, F. Yang, M. Lv, Z. Li and X. Liu, J. Mater. Chem. A, 2023, 11, 5476–5494 RSC.
  2. S. A. Thomas and J. Cherusseri, J. Mater. Chem. A, 2023, 11, 23148–23187 RSC.
  3. G. M. Tomboc, J. Kim, Y. Wang, Y. Son, J. Li, J. Y. Kim and K. Lee, J. Mater. Chem. A, 2021, 9, 4528–4557 RSC.
  4. H. Lee, J. Park, J.-M. Myoung and B. Han, J. Mater. Chem. A, 2025,(13), 9244–9251 RSC.
  5. S. Balachandran, K. Jeeva Jothi, K. Selvakumar, D. K. Bhat, K. Sathiyanarayanan and M. Swaminathan, Mater. Chem. Phys., 2020, 256, 123624 CrossRef CAS.
  6. A. De, M. S. Kim, A. Adhikari, R. Patel and S. Kundu, J. Mater. Chem. A, 2024, 12, 19720–19756 RSC.
  7. J. Libich, J. Máca, J. Vondrák, O. Čech and M. Sedlaříková, J. Energy Storage, 2018, 17, 224–227 CrossRef.
  8. A. Shiralizadeh Dezfuli, E. Kohan, H. R. Naderi and E. Salehi, New J. Chem., 2019, 43, 9260–9264 RSC.
  9. W. Sun and Z. Mo, Mater. Sci. Eng., B, 2013, 178, 527–532 CrossRef CAS.
  10. R. Zou, H. Quan, M. Pan, S. Zhou, D. Chen and X. Luo, Electrochim. Acta, 2018, 292, 31–38 CrossRef CAS.
  11. M.-K. Song, S. Cheng, H. Chen, W. Qin, K.-W. Nam, S. Xu, X.-Q. Yang, A. Bongiorno, J. Lee, J. Bai, T. A. Tyson, J. Cho and M. Liu, Nano Lett., 2012, 12, 3483–3490 CrossRef CAS PubMed.
  12. W. Chen, R. B. Rakhi, L. Hu, X. Xie, Y. Cui and H. N. Alshareef, Nano Lett., 2011, 11, 5165–5172 CrossRef CAS PubMed.
  13. R. Liu, L. Ma, G. Niu, X. Li, E. Li, Y. Bai and G. Yuan, Adv. Funct. Mater., 2017, 27, 1701635 CrossRef.
  14. Y. Zhu, X. Liu, S. Jin, H. Chen, W. Lee, M. Liu and Y. Chen, J. Mater. Chem. A, 2019, 7, 5875–5897 RSC.
  15. D. Gogoi, R. S. Karmur, M. R. Das and N. N. Ghosh, J. Mater. Chem. A, 2023, 11, 23867–23880 RSC.
  16. C. Yang, Y. Lu, W. Duan, Z. Kong, Z. Huang, T. Yang, Y. Zou, R. Chen and S. Wang, Energy Fuels, 2021, 35, 14283–14303 CrossRef CAS.
  17. A. Bashir, A. Inayat, R. Bashir, S. Jamil, S. M. Abbas, M. Sultan, A. Iqbal and Z. Akhter, New J. Chem., 2023, 47, 3560–3571 RSC.
  18. Q. Yue, J. Sun, S. Chen, Y. Zhou, H. Li, Y. Chen, R. Zhang, G. Wei and Y. Kang, ACS Appl. Mater. Interfaces, 2020, 12, 18570–18577 CrossRef CAS PubMed.
  19. D. Ontiveros, F. Viñes and C. Sousa, J. Mater. Chem. A, 2023, 11, 13754–13764 RSC.
  20. S. Ponnada, M. S. Kiai, O. Eroglu, C.-O. Kim, R. K. Sharma and G. S. Martynkova, Energy Fuels, 2024, 38, 18242–18264 CrossRef CAS.
  21. A. Raveendran, M. Chandran, M. R. Siddiqui, S. M. Wabaidur, M. Eswaran and R. Dhanusuraman, ACS Omega, 2023, 8, 34768–34786 CrossRef CAS PubMed.
  22. K. Yang, H. Cheng, B. Wang, Y. Tan, T. Ye, Y. Yang and C. Wang, Adv. Mater. Technol., 2022, 7, 2100675 CrossRef CAS.
  23. S. Venkateshalu, G. M. Tomboc, S. P. Nagalingam, J. Kim, T. Sawaira, K. Sehar, B. G. Pollet, J. Y. Kim, A. Nirmala Grace and K. Lee, J. Mater. Chem. A, 2023, 11(27), 14469–14488 RSC.
  24. S. N. Ansari, M. Saraf, Z. Abbas and S. M. Mobin, Nanoscale, 2023, 15, 13546–13560 RSC.
  25. X. Liu, M. Wang, B. Qin, Y. Zhang, Z. Liu and H. Fan, Chem. Eng. J., 2022, 431, 133796 CrossRef CAS.
  26. J. B. Lee, G. H. Choi and P. J. Yoo, J. Alloys Compd., 2021, 887, 161304 CrossRef CAS.
  27. Q. Zhang, X. Wang, Z. Pan, J. Sun, J. Zhao, J. Zhang, C. Zhang, L. Tang, J. Luo, B. Song, Z. Zhang, W. Lu, Q. Li, Y. Zhang and Y. Yao, Nano Lett., 2017, 17, 2719–2726 CrossRef CAS PubMed.
  28. M. Majumder, R. B. Choudhary, A. K. Thakur and U. Kumar, J. Electroanal. Chem., 2017, 804, 42–52 CrossRef CAS.
  29. R. Tong, Z. Sun, X. Wang, S. Wang and H. Pan, ChemElectroChem, 2019, 6, 1338–1343 CrossRef CAS.
  30. Q. Ke, C. Guan, X. Zhang, M. Zheng, Y.-W. Zhang, Y. Cai, H. Zhang and J. Wang, Adv. Mater., 2017, 29, 1604164 CrossRef PubMed.
  31. A. Padhy, S. Lakshmy, B. Chakraborty and J. N. Behera, Sustainable Energy Fuels, 2023, 7, 5271–5282 RSC.
  32. S. Yang, Y. Liu, Y. Hao, X. Yang, W. A. Goddard III, X. L. Zhang and B. Cao, Adv. Sci., 2018, 5, 1700659 CrossRef PubMed.
  33. M. Z. Iqbal, M. Shaheen, U. Aftab, Z. Ahmad, M. Y. Solangi, M. I. Abro and S. M. Wabaidur, Energy Fuels, 2024, 38, 2416–2425 CrossRef CAS.
  34. M. M. Ibrahim, A. Mezni, M. Alsawat, T. Kumeria, M. R. Das, S. Alzahly, A. Aldalbahi, K. Gornicka, J. Ryl, M. A. Amin and T. Altalhi, Int. J. Energy Res., 2019, 43, 5367–5383 CrossRef CAS.
  35. M. Zhi, C. Xiang, J. Li, M. Li and N. Wu, Nanoscale, 2013, 5, 72–88 RSC.
  36. C. An, Y. Zhang, H. Guo and Y. Wang, Nanoscale Adv., 2019, 1, 4644–4658 RSC.
  37. Y. Ma, X. Xie, W. Yang, Z. Yu, X. Sun, Y. Zhang, X. Yang, H. Kimura, C. Hou, Z. Guo and W. Du, Adv. Compos. Hybrid Mater., 2021, 4, 906–924 CrossRef CAS.
  38. J. Liu, W. Chen, Q. Sun, Y. Zhang, X. Li, J. Wang, C. Wang, Y. Yu, L. Wang and X. Yu, ACS Appl. Energy Mater., 2021, 4, 2864–2872 CrossRef CAS.
  39. Y. He, W. Zhou, D. Li, Y. Liang, S. Chao, X. Zhao, M. Zhang and J. Xu, Small, 2023, 19, 2206956 CrossRef CAS PubMed.
  40. D. Yan, W. Wang, X. Luo, C. Chen, Y. Zeng and Z. Zhu, Chem. Eng. J., 2018, 334, 864–872 CrossRef CAS.
  41. J. Hao, S. Peng, H. Li, S. Dang, T. Qin, Y. Wen, J. Huang, F. Ma, D. Gao, F. Li and G. Cao, J. Mater. Chem. A, 2018, 6, 16094–16100 RSC.
  42. X. Zhao, Y. Jing, Z. Dai, Y. Chu, Z. Liu, Y. Cong and J. Song, ACS Omega, 2024, 9, 16868–16875 CrossRef CAS PubMed.
  43. F. Lei, Y. Sun, K. Liu, S. Gao, L. Liang, B. Pan and Y. Xie, J. Am. Chem. Soc., 2014, 136, 6826–6829 CrossRef CAS PubMed.
  44. A. Zhang, R. Gao, L. Hu, X. Zang, R. Yang, S. Wang, S. Yao, Z. Yang, H. Hao and Y.-M. Yan, Chem. Eng. J., 2021, 417, 129186 CrossRef CAS.
  45. M. Sun, R.-T. Gao, J. He, X. Liu, T. Nakajima, X. Zhang and L. Wang, Angew. Chem., Int. Ed., 2021, 60, 17601–17607 CrossRef CAS PubMed.
  46. Q. Yang, Y. Wang, Q. Tian, X. Li, A. Pan, M. Zhao, Y. Zhu, T. Wu and G. Fang, J. Mater. Chem. A, 2024, 12, 7207–7214 RSC.
  47. P. Aryanrad, H. R. Naderi, E. Kohan, M. R. Ganjali, M. Baghernejad and A. Shiralizadeh Dezfuli, RSC Adv., 2020, 10, 17543–17551 RSC.
  48. V. Sharma, R. Dhiman and A. Mahajan, Appl. Surf. Sci., 2023, 612, 155883 CrossRef CAS.
  49. J. D. Benck, Z. Chen, L. Y. Kuritzky, A. J. Forman and T. F. Jaramillo, ACS Catal., 2012, 2, 1916–1923 CrossRef CAS.
  50. M. H. Suliman, A. Adam, M. N. Siddiqui, Z. H. Yamani and M. Qamar, Carbon, 2019, 144, 764–771 CrossRef CAS.
  51. T. Li and D. Ding, Materials, 2019, 12(24), 4102 CrossRef CAS PubMed.
  52. S. Sardana, P. Mahajan, A. Mishra, A. K. Chawla and A. Mahajan, Adv. Mater. Technol., 2024, 2400829 Search PubMed.
  53. A. Padhy, R. Samal, C. S. Rout and J. N. Behera, Sustainable Energy Fuels, 2022, 6, 2010–2019 RSC.
  54. P. Hohenberg and W. Kohn, Phys. Rev. [Sect.] B, 1964, 136, B864–B871 CrossRef.
  55. P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari and R. M. Wentzcovitch, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
  56. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  57. J. Singh, G. P. Singh, K. J. Singh, N. Virdi, D. A. Tonpe and R. C. Singh, Desalin. Water Treat., 2025, 322, 101155 CrossRef.
  58. D. Cao, H. Liu, Q. Sun, X. Wang, J. Wang, C. Liu, C. Liu, Q. Hao and Y. Li, Appl. Surf. Sci., 2021, 563, 150210 CrossRef CAS.
  59. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef.
  60. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  61. C. G. Broyden, IMA J. Appl. Math., 1970, 6, 222–231 CrossRef.
  62. R. Fletcher, Comput. J., 1970, 13, 317–322 CrossRef.
  63. Z. Wang, H. Di, R. Sun, Y. Zhu, L. Yin, Z. Zhang and C. Wang, ChemPhysMater, 2022, 1, 321–329 CrossRef.
  64. S. Gasso, M. K. Sohal and A. Mahajan, Sens. Actuators, B, 2022, 357, 131427 CrossRef CAS.
  65. H. Wang, R. Zhao, J. Qin, H. Hu, X. Fan, X. Cao and D. Wang, ACS Appl. Mater. Interfaces, 2019, 11, 44249–44262 CrossRef CAS PubMed.
  66. T. R. Poorani, C. Ramya and M. Ramya, J. Inorg. Organomet. Polym. Mater., 2023, 33, 3413–3440 CrossRef CAS.
  67. B. Sun, Y. Qian, Z. Liang, Y. Guo, Y. Xue, J. Tian and H. Cui, Sol. Energy Mater. Sol. Cells, 2019, 195, 309–317 CrossRef CAS.
  68. C.-H. Zeng, K. Zheng, K.-L. Lou, X.-T. Meng, Z.-Q. Yan, Z.-N. Ye, R.-R. Su and S. Zhong, Electrochim. Acta, 2015, 165, 396–401 CrossRef CAS.
  69. K. L. Firestein, J. E. von Treifeldt, D. G. Kvashnin, J. F. S. Fernando, C. Zhang, A. G. Kvashnin, E. V Podryabinkin, A. V Shapeev, D. P. Siriwardena, P. B. Sorokin and D. Golberg, Nano Lett., 2020, 20, 5900–5908 CrossRef CAS PubMed.
  70. J. Yang, Z. Quan, D. Kong, X. Liu and J. Lin, Cryst. Growth Des., 2007, 7, 730–735 CrossRef CAS.
  71. V. Natu, M. Benchakar, C. Canaff, A. Habrioux, S. Célérier and M. W. Barsoum, Matter, 2021, 4, 1224–1251 CrossRef CAS.
  72. A. Karmakar, K. Karthick, S. S. Sankar, S. Kumaravel, R. Madhu, K. Bera, H. N. Dhandapani, S. Nagappan, P. Murugan and S. Kundu, J. Mater. Chem. A, 2022, 10, 3618–3632 RSC.
  73. Y. Zare, Composites, Part A, 2016, 84, 158–164 CrossRef CAS.
  74. S. Kumar, R. Prakash, R. J. Choudhary and D. M. Phase, Mater. Res. Bull., 2015, 70, 392–396 CrossRef CAS.
  75. T. Zhang, M.-Y. Wu, D.-Y. Yan, J. Mao, H. Liu, W.-B. Hu, X.-W. Du, T. Ling and S.-Z. Qiao, Nano Energy, 2018, 43, 103–109 CrossRef CAS.
  76. Y. Guo, Y. Tong, P. Chen, K. Xu, J. Zhao, Y. Lin, W. Chu, Z. Peng, C. Wu and Y. Xie, Adv. Mater., 2015, 27, 5989–5994 CrossRef CAS PubMed.
  77. G. Li, R. Huang, C. Zhu, G. Jia, S. Zhang and Q. Zhong, Colloids Surf., A, 2021, 614, 126156 CrossRef CAS.
  78. S. P. Narayanan, P. Thakur, A. P. Balan, A. A. Abraham, F. Mathew, M. Yeddala, T. Subair, C. Tiwary, S. Thomas, T. N. Narayanan, P. M. Ajayan and M. M. R. Anantharaman, Phys. Status Solidi RRL, 2019, 13, 1900025 CrossRef CAS.
  79. S. A. Ali and T. Ahmad, Int. J. Hydrogen Energy, 2023, 48, 22044–22059 CrossRef CAS.
  80. A. Ahmed, M. Naseem Siddique, U. Alam, T. Ali and P. Tripathi, Appl. Surf. Sci., 2019, 463, 976–985 CrossRef CAS.
  81. X. Zhang, X. Liu, Y. Zeng, Y. Tong and X. Lu, Small Methods, 2020, 4, 1900823 CrossRef CAS.
  82. G. Wang and Z. Jin, J. Mater. Chem. C, 2021, 9, 620–632 RSC.
  83. C. Kang, C. Li, J. Zhang, J. Ren and M. Chen, Int. J. Hydrogen Energy, 2024, 78, 503–512 CrossRef CAS.
  84. Y. Su, M. Song, X. Wang, J. Jiang, X. Si, T. Zhao and P. Qian, Nanomaterials, 2021, 11(10), 2497 CrossRef CAS PubMed.
  85. I. Ullah, A. Irfan, M. Li, S. Saddique, T. Yang, H. Zhao, N. Irshad, K. Yang, C. Wang and P. K. Chu, J. Power Sources, 2025, 643, 237046 CrossRef CAS.
  86. M. Das and S. Ghosh, J. Electrochem. Soc., 2022, 169, 90525 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta01195g

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.