Next Article in Journal
Hydrothermal Aging Mechanism of CeO2-Based Catalytic Materials and Its Structure–Activity Relationship Study on Particulate Matter Oxidation Performance
Previous Article in Journal
Enhancement of Photocatalytic and Anticancer Properties in Y2O3 Nanocomposites Embedded in Reduced Graphene Oxide and Carbon Nanotubes
Previous Article in Special Issue
The Role of the Transition Metal in M2P (M = Fe, Co, Ni) Phosphides for Methane Activation and C–C Coupling Selectivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Integration of TiO2 Nanorods with Carbon Cloth for Enhanced Photocatalytic Hydrogen Evolution and Wastewater Remediation

by
Shakeelur Raheman AR
1,2,
Khursheed B. Ansari
3,*,
Sang Joon Lee
1,* and
Nilesh Salunke
4
1
Department of Mechanical Engineering, Pohang University of Science and Technology, 77 Cheongam-Ro, Nam-Gu, Pohang 37673, Gyeongbuk, Republic of Korea
2
Department of Applied Sciences & Humanities, Shri Vile Parle Kelavani Mandal’s IoT Dhule, Dhule 424001, Maharashtra, India
3
Department of Chemical Engineering, College of Engineering, King Khalid University, Abha 61411, Saudi Arabia
4
Department of Mechanical Engineering, Shri Vile Parle Kelavani Mandal’s IoT Dhule, Dhule 424001, Maharashtra, India
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(10), 961; https://doi.org/10.3390/catal15100961
Submission received: 8 September 2025 / Revised: 26 September 2025 / Accepted: 4 October 2025 / Published: 7 October 2025
(This article belongs to the Special Issue Advanced Catalysis for Energy and a Sustainable Environment)

Abstract

The immobilization of titanium dioxide (TiO2) nanostructures on conductive supports offers a promising strategy to overcome the intrinsic limitations of a wide band gap, poor visible-light absorption, and rapid charge recombination in photocatalysis. Herein, a rutile TiO2 nanorods (TiO2NRs) array was directly grown on carbon cloth (CC) via a hydrothermal method by using titanium tetrachloride (TiCl4) seed solutions of 0.1, 0.3, and 0.5 M, designated as TiO2NR0.1/CC, TiO2NR0.3/CC, and TiO2NR0.5/CC, respectively. Structural analysis confirmed that the TiO2 NRs array is vertically aligned, and phase=pure rutile NRs strongly adhered to CC. The optical characterization revealed broadened absorption in the visible wavelength region and progressive band gap narrowing with the increasing seeding concentration. Photoluminescence (PL) spectra showed pronounced quenching in the fabricated TiO2NRs/CC samples, especially with TiO2NR0.3/CC exhibiting the lowest PL intensity, indicating suppressed charge recombination. Electrochemical impedance spectroscopy further demonstrated reduced charge transfer resistance, and TiO2NR0.3/CC achieved the most efficient electron transport kinetics. Photocatalytic tests at λ ≥ 400 nm irradiation confirmed the enhanced hydrogen evolution performance of TiO2NR0.3/CC. The hydrogen yield of 2.66 mmol h−1 g−1 of TiO2NR0.3/CC was 4.03-fold higher than that of TiO2NRs (0.66 mmol h−1 g−1), along with excellent cyclic stability across three runs. Additionally, TiO2NR0.3/CC achieved 90.2% degradation of methylene blue within 60 min, with a kinetic constant of 0.0332 min−1 and minimal activity loss after three cycles. These results highlight the synergistic integration of TiO2 NRs with CC in achieving a durable, recyclable, and efficient photocatalytic platform for sustainable hydrogen generation and wastewater remediation.

1. Introduction

The continuous exhaustion of fossil fuel reserves and rising concerns over environmental pollution have accelerated the search for sustainable and renewable solutions for energy generation and wastewater remediation. Among emerging technologies, photocatalysis has attracted large attention as a green strategy that utilizes solar energy to drive both hydrogen (H2) production through water splitting and organic pollutants degradation in wastewater [1,2]. Hydrogen, as a clean fuel with high energy density, represents an important element for future energy systems, while the removal of toxic dyes from textile and chemical industries is crucial to mitigating ecological risks. However, the effectiveness of photocatalytic treatment strongly depends on the design of efficient, durable, and recyclable photocatalysts, capable of overcoming intrinsic material limitations. In this context, TiO2 has remained one of the most extensively studied semiconductor photocatalysts due to its chemical stability, low toxicity, abundance, and photocorrosion resistance. One-dimensional TiO2 nanostructures, i.e., nanorods (NRs) and nanowires (NWRs), have gained particular interest because they provide oriented electron pathways, high aspect ratios, and improved light-harvesting ability compared to nanoparticulate TiO2 [3]. Guo et al. reported the growth of rectangular bunched rutile TiO2 NRs arrays on CC, enhancing the photocatalytic performance compared to conventional nanoparticle films [4]. Nevertheless, pristine TiO2 often suffers from a wide band gap (~3.0–3.2 eV), restricting its activity primarily in the ultraviolet wavelength region that constitutes ~5% of the solar spectrum, and rapid recombination of photogenerated electron–hole pairs limits quantum efficiency [5]. Moreover, when employed as powder suspensions, TiO2 catalysts pose difficulties in recovery and recycling, which hinders their large-scale deployment in wastewater treatment and H2 evolution systems [5,6]. To address these challenges, some researchers explored the immobilization of photocatalysts on conductive and flexible supports. Carbon cloth (CC), a woven textile composed of carbon microfibers, has recently emerged as an attractive substrate for photocatalysis. Its high electrical conductivity allows the rapid extraction and transfer of photogenerated electrons from semiconductors, thereby suppressing electron–hole recombination and improving catalytic efficiency [7,8]. The mechanical flexibility and chemical stability of CC enable the fabrication of bendable, durable photocatalytic devices suitable for integration into flow systems or flexible reactors [9]. Furthermore, the three-dimensional porous network of CC facilitates the mass transfer of reactants and products, while its large surface area supports dense and uniform loading of nanostructured catalysts [10,11]. Importantly, the immobilization of photocatalysts on CC eliminates the need for centrifugation or filtration after reactions, enabling easy recyclability and preventing secondary pollution [12]. These features make CC a multifunctional scaffold that enhances photocatalytic activity, while ensuring practical applicability.
The advantages of CC have been demonstrated in numerous nanomaterials immobilized on CC for environmental and energy-related photocatalysis. Yi et al. synthesized ZnO@Ag3PO4 core–shell structures on CC, and the archived rhodamine B degradation efficiency was nearly 5–16 times higher than individual components due to efficient interfacial charge separation and fast electron transfer through CC [13]. Jian et al. reported that Co3O4/AgIO4 heterojunctions on CC showed approximately 58-fold enhancement in dye degradation compared to bare Co3O4. This was attributed to synergistic S-scheme charge transfer mediated by conductive carbon fibers [14]. Chang et al. fabricated BiOBr/CuO composites on CC, which functioned as immobilized photocatalytic membranes with remarkable H2 evolution activity under visible light [10]. Jiang et al. designed oxygen-deficient WO3/TiO2 NRs on CC, in which the conductive substrate collectively improved charge separation and visible-light absorption [12]. In another study, zinc indium sulfide (ZnIn2S4) nanosheets grown on carbon fibers delivered enhanced photocatalytic H2 evolution, confirming that intimate coupling with CC facilitates directional electron flow [15,16]. These reports collectively highlight that CC not only serves as a passive support but actively contributes to improved photocatalytic charge dynamics.
Compared with powder catalysts, CC-supported systems exhibit multiple scientific advantages such as (i) electron transfer enhancement due to excellent conductivity of carbon fibers, which act as electron reservoirs, (ii) improved light utilization, as the woven 3D texture allows multidirectional photon penetration and scattering, (iii) structural stability and flexibility, enabling durable operation under mechanical stress and continuous operation of flow reactors, (iv) recyclability, as CC-supported photocatalysts can be easily retrieved and reused, and (v) interfacial synergy, where strong contact between nanomaterials and CC suppresses recombination and prolongs carrier lifetimes [16]. Despite these promising advantages of CC supporting various nanomaterials, relatively few studies have systematically investigated TiO2 NR arrays grown directly on CC for dual photocatalytic functions of H2 evolution and dye degradation. The aligned TiO2 NRs’ morphology provides one-dimensional (1D) pathways for electron transport, while the CC substrate extracts electrons efficiently, thereby suppressing recombination and enhancing both reduction (H2 generation) and oxidation (dye degradation) reactions. Moreover, the immobilization on CC overcomes the recovery issues, allowing for long-term operational stability and practical reusability. However, the detailed understanding of how nanorod growth orientation, interface structure, and defect states influence charge separation and photocatalytic efficiency in TiO2NRs/CC systems remains limited.
Therefore, the present work focuses on the rational design and fabrication of TiO2 NRs arrays directly grown on CC, and evaluates their photocatalytic performance for H2 evolution and organic dye degradation under solar irradiation conditions. The present system uniquely achieves solar-to-hydrogen conversion without any costly noble metal co-catalysts, as well as exhibiting simultaneous organic dye degradation with good recyclability across multiple cycles.
By leveraging the synergistic integration of 1D TiO2 with the conductive and flexible carbon substrate, this study aims to (i) enhance charge carrier separation and electron transport, (ii) maximize surface-active sites for photocatalytic reactions, (iii) ensure recyclability and structural durability, and (iv) provide insights into the mechanistic role of CC in promoting photocatalytic efficiency. This approach not only advances the application of TiO2-based systems but also contributes to the new development of flexible, multifunctional photocatalytic platforms for addressing global energy and environmental challenges.

2. Results and Discussion

2.1. Surface Structure of TiO2 NRs/CC

The XRD patterns of pristine CC and TiO2 NRs grown on CC (i.e., TiO2NRs0.3/CC) are shown in Figure 1. The pristine CC exhibits a broad hump at around 25°, corresponding to the amorphous graphitic carbon phase [17]. In contrast, the TiO2NRs/CC exhibits sharp diffraction peaks located at 2θ = 27.88°, 36.60°, 39.62°, 41.76°, 44.52°, 54.76°, and 57.02°, which can be indexed to the (110), (101), (200), (111), (210), (211), and (220) planes of rutile TiO2. The absence of characteristic anatase (101) reflection at 25.3° or brookite-related peaks confirms the formation of phase-pure rutile TiO2 by hydrothermal treatment [18]. The dominant intensity of the (110) reflection indicates preferential orientation along this plane, which is a typical feature of rutile TiO2 NRs grown in strongly acidic chloride-containing media. This result is consistent with earlier reports that chloride ions selectively adsorbed onto certain rutile crystal facets, thereby suppressing lateral growth and promoting anisotropic elongation along the c-axis to form rod-like structures [19]. The sharp and well-defined peaks further demonstrate the high crystallinity of the obtained TiO2 NRs. This high crystallinity is beneficial for efficient charge transport in photocatalytic and photoelectrochemical applications [20]. The combination of seeding with TiCl4 and subsequent hydrothermal growth ensured uniform nucleation and dense coverage of TiO2 NRs on the CC substrate. This high crystallinity and phase purity might be helpful for enhancing electron transport and photocatalytic activity, making the TiO2 NR/CC composites suitable for photoelectrochemical and energy-related applications [21].

2.2. Surface Morphology of TiO2 NRs/CC

The morphological features of the pristine CC and TiO2 NRs grown on CC were examined by employing scanning electron microscopy (SEM) and transmission electron microscopy (TEM) analyses. Figure 2a,b shows the surface morphologies of pristine CC, where the individual carbon fibers have exhibited a smooth surface, and they have a tightly woven textile structure. The absence of surface deposits confirms that the acid pretreatment effectively removed impurities, leaving behind a clean substrate suitable for TiO2 nucleation. Upon hydrothermal growth, significant structural modifications were observed. The digital photograph in Figure 2c shows the overall whitish appearance of TiO2NR0.3/CC, indicating uniform deposition of TiO2 nanostructures. At lower SEM magnification (cf. Figure 2d), the carbon fibers are seen to be uniformly coated with vertically aligned TiO2 NRs, forming a dense nanorod network across the substrate. With higher magnification (i.e., Figure 2e,f), it becomes clear that the nanorods are radially anchored onto the carbon fiber surface, creating a roughened texture compared to pristine CC. The nanorods are densely packed with diameters in the tens of nanometers and lengths extending into the micrometer range. Such dense vertical alignment can be attributed to the TiCl4 seeding and chloride-assisted hydrothermal conditions, which promote anisotropic rutile TiO2 growth through selective facet stabilization [20,22]. The TEM micrographs provide further insights into the structural features. Figure 2g shows a cluster of elongated TiO2 NRs with uniform size distribution, while the high-resolution TEM image in Figure 2h displays individual nanorods with clear lattice fringes, confirming their single-crystalline nature. The SAED pattern in Figure 2i exhibits sharp diffraction spots characteristic of a well-ordered rutile crystal structure, in agreement with the XRD analysis. The formation of single-crystalline rutile nanorods directly on a conductive and flexible CC substrate ensures strong interfacial adhesion, enhanced electron transport, and high surface area, all of which are highly desirable for the practical applications of photocatalysis.

2.3. Surface Chemical Composition and States of TiO2 NRs/CC

The chemical composition and oxidation states of TiO2NR0.3/CC were examined by X-ray photoelectron spectroscopy (XPS). The survey spectrum (cf. Figure 3a) confirmed the presence of Ti, O, and C elements on the surface. The high-resolution Ti 2p spectrum (cf. Figure 3b) shows two peaks at ~458.6 eV (Ti 2p) and ~463.3 eV (Ti 2p1/2), with a spin–orbit splitting of ~5.7 eV. These values are characteristic of Ti4+ in stoichiometric TiO2, confirming that titanium exists predominantly in the fully oxidized state without evidence of Ti3+ [23,24]. The O 1s spectrum (cf. Figure 3c) can be fitted into two components. The main peak at ~530.1 eV is attributed to lattice oxygen (Ti–O–Ti), while the higher binding energy shoulder at ~531.6 eV corresponds to surface hydroxyl groups or adsorbed oxygen species [25]. The presence of hydroxyl groups is advantageous, as they promote surface reactivity and improve photocatalytic performance. The C 1s spectrum exhibits a dominant peak at ~284.8 eV due to graphitic sp2 carbon from the carbon fiber substrate, while the minor peaks at ~286.5–288.5 eV are associated with C–O and O–C=O groups (see Figure 3d). Such oxygen-containing functionalities are known to arise from oxidative pretreatment and serve as anchoring sites for TiO2 nucleation. This analysis confirms the successful growth of rutile TiO2 NRs with a Ti4+ oxidation state, strong bonding to the carbon fiber substrate, and surface hydroxylation, all of which are favorable for photocatalytic applications.

2.4. Optical and Photophysical Properties of TiO2 NRs/CC

The optical and photophysical properties of TiO2NRs/CC were systematically examined through UV–Vis absorbance (Figure 4a,b), Tauc band gap analysis (Figure 4c), and photoluminescence spectroscopy (Figure 4d–f). Figure 5a shows the diffuse reflectance UV–Vis spectra, revealing broadened absorption extending from the UV into the visible region, particularly for samples seeded with higher TiCl4 concentrations. This red shift and enhanced visible-light absorption align with previous findings that carbonaceous supports effectively extend TiO2’s light-harvesting capabilities through improved interfacial interactions and charge transfer dynamics. In Figure 4c, pristine TiO2 NRs exhibited a band gap of 3.08 eV, corresponding to an absorption edge near 403 nm. When integrated with carbon cloth, a slight reduction in the band gap was observed, with values of 3.01 eV, 2.99 eV, and 2.99 eV for TiO2 NRs grown using 0.1 M, 0.3 M, and 0.5 M TiCl4, respectively. This red shift of about 0.07–0.09 eV indicates improved light absorption due to the interaction between TiO2 and the conductive carbon substrate. The shift is likely attributed to interfacial states, defect levels, or C–Ti–O bonding, which extend the absorption edge marginally into the near-UV–visible region. Importantly, increasing the TiCl4 concentration from 0.3 M to 0.5 M did not significantly alter the band gap, suggesting that the optical modification saturates around 0.3 M. This implies that further increases in the precursor concentration primarily influence the nanorod density and morphology rather than the electronic structure. From a practical perspective, 0.3 M TiCl4 offers an optimal balance, as it achieves the desired band gap narrowing while minimizing excess precursor use. This trend is consistent with the literature reports where carbon doping or carbon support leads to band gap narrowing in TiO2 composites [26]. The photoluminescence (PL) spectra of TiO2 NRs deposited on CC are deconvoluted, as shown in Figure 4d–f. Two characteristic emission bands are observed for all samples, with the first peak appearing at 437 nm for TiO2NRs0.1/CC and ~438 nm for TiO2NRs0.3/CC and TiO2NRs0.5/CC, corresponding to near band-edge excitonic recombination. The second peak occurs around 465–466 nm and is attributed to intrinsic defect states, mainly oxygen vacancies and Ti3+ centers. The maximum PL intensities were found to be 592 a.u. for TiO2NRs0.1/CC, 462 a.u. TiO2NRs0.3/CC, and 559 a.u. for TiO2NRs0.5/CC. Especially, TiO2NRs0.3/CC shows the strongest quenching effect. This reduction in PL intensity demonstrates improved charge separation and longer carrier lifetimes due to the conductive CC acting as an electron sink. Such enhanced charge dynamics are essential for maximizing H2 generation efficiency and ensuring effective degradation of organic pollutants. Interestingly, at the 0.5 M concentration, the loading in PL intensity is slightly increased. This seems to be attributed to excess CC hindering light absorption or introducing recombination centers [27].

2.5. Charge Transport Resistance and Semiconductor Characteristics of TiO2 NRs/CC

Electrochemical impedance spectroscopy (EIS) and Mott–Schottky (M–S) analyses were further employed to probe the charge transport resistance and semiconductor characteristics of the TiO2NRs/CC electrodes (Figure 5a–c). The Nyquist plots in Figure 6a represent that the bare TiO2 NRs exhibit the largest semicircle diameter, corresponding to a higher charge transfer resistance (Rct), whereas the TiO2NRs/CC composites display substantially reduced semicircles, confirming faster interfacial charge transport. Among them, TiO2NRs0.3/CC demonstrates the smallest semicircle, indicating the most efficient electron transfer kinetics at the electrode–electrolyte interface. This improvement is attributed to the synergistic role of the CC, which enhances electrical conductivity and suppresses recombination by providing a direct pathway for electron collection. Mott–Schottky measurements (Figure 5b and c) further support these findings by elucidating the flat-band potential and carrier density. All potentials initially measured against the Ag/AgCl reference electrode were converted to the RHE scale by using the Nernst equation (ERHE = EAg/AgCl + 0.197 + 0.059 × pH), where 0.197 V represents the standard potential of Ag/AgCl (saturated KCl) at 25 °C [2]. For pristine TiO2 NRs (Figure 5b), the flat-band potential is observed at approximately –0.2 V vs. RHE, whereas for TiO2NRs0.3/CC (Figure 5c), a more negative shift to –0.65 V is recorded. This cathodic shift suggests that CC incorporation not only alters the electronic environment but also enhances band bending, facilitating charge separation and a stronger driving force for reduction reactions [28]. Moreover, the slope of TiO2NRs/CC electrodes implies a higher donor density compared to pure TiO2, which is consistent with the literature reports on carbon-supported TiO2 photoelectrodes. These electrochemical results collectively confirm that an optimal TiCl4 concentration (0.3/CC) yields the most favorable charge transport and electronic properties. Excessive carbon loading (0.5/CC), however, slightly deteriorates performance. This might result from the partial blocking of active TiO2 sites or the introduction of recombination centers, similar to previous observations in TiO2/carbon hybrid systems [29]. Linear sweep voltammetry (LSV) measurements were further carried out to evaluate the photoelectrochemical (PEC) activity of the TiO2NRs/CC electrodes (cf. Figure 5d). The experiments were performed in a standard three-electrode configuration using 0.5 M Na2SO4 aqueous solution (pH ~7) as the electrolyte under light illumination (AM 1.5G, 100 mW cm−2). The bare CC electrode exhibited a very low photocurrent density (~2 × 10−4 mA cm−2 at 1.0 V vs. RHE), highlighting its limited contribution to photoactivity. Upon the integration of TiO2 nanorods, the photocurrent density increased significantly, confirming the beneficial role of TiO2 in light harvesting and charge generation. Notably, the TiO2NRs0.3/CC electrode delivered the highest photocurrent density of ~2.1 × 10−3 mA cm−2, outperforming both TiO2NRs0.1/CC (~1.0 × 10−3 mA cm−2) and TiO2NRs0.5/CC (~1.5 × 10−3 mA cm−2). This enhanced performance is attributed to the optimized nanorod morphology and interfacial contact achieved at the 0.3 M TiCl4 precursor concentration, which synergistically promotes efficient charge separation and transport while maintaining a large active surface area. In contrast, an excessive Ti precursor concentration (0.5 M) likely leads to the overgrowth and partial aggregation of nanorods, which hinder charge transfer and increase recombination losses, thereby reducing the photocurrent output. These LSV results, in agreement with EIS and M–S analyses, demonstrate that 0.3 M TiCl4 is the optimal condition for fabricating TiO2NRs/CC electrodes with superior PEC performance.

2.6. Photocatalytic Performance of TiO2NRs/CC

2.6.1. H2 Evolution

The TiO2 NRs-coated carbon fabric was placed on a polyurethane foam ring to float on an electrolyte surface in a splitable borosilicate glass reactor containing 20 mL of aqueous solution (90 vol% H2O, 10 vol% triethanolamine (TEOA)). The reactor was purged with nitrogen gas for 30–60 min to remove oxygen. The sealed reactor was exposed to 100 mW/cm2 light irradiation, and the reaction temperature was maintained at 25 °C using a cooling system. The photocatalytic H2 evolution reaction was conducted for 1–4 h, with periodic headspace gas sampling to quantify H2 production via GC analysis. Figure 6a and Figure 6b show the H2 evolution and rate of H2 evolution, respectively, using the as-prepared photocatalysts. Here, the bare TiO2 NRs showed relatively low H2 generation, producing only 41.8 µmol of H2 after 4 h of illumination with a rate of 0.66 mmol h−1 g−1. When TiO2 NRs were deposited on CC, the photocatalytic activity improved significantly due to enhanced electron transfer and reduced recombination. At 0.1 M TiCl4, the composite produced 68.2 µmol of H2 with a rate of 0.90 mmol h−1 g−1, while further increasing the concentration to 0.3 M resulted in a remarkable enhancement, reaching 120.5 µmol of H2 after 4 h with the highest H2 evolution rate of 2.66 mmol h−1 g−1. At 0.5 M TiCl4, the H2 evolution was slightly lower, with 114.3 µmol after 4 h and a rate of 2.24 mmol h−1 g−1, which suggests that an excessive precursor concentration may lead to denser growth and hinder charge mobility. Figure 6c shows the cyclic stability tests conducted for the optimized TiO2NRs0.3M/CC sample showed excellent reproducibility across three consecutive cycles, with H2 evolution values of 122.5, 120.5, and 121.5 µmol after 4 h, indicating negligible loss in activity and confirming the robustness of the photocatalyst. These results demonstrate that the combination of TiO2 NRs with a conductive CC substrate, particularly when synthesized with 0.3 M TiCl4, provides an efficient and durable system for photocatalytic H2 production in TEOS aqueous solution.

2.6.2. Dye Degradation

The photocatalytic degradation of methyl blue (MB) dye using TiO2NRs/CC is investigated under visible-light irradiation, as shown in Figure 7a–d. The UV–Vis absorption spectra (Figure 7a) clearly demonstrate the gradual reduction in MB intensity with the increasing irradiation time, confirming continuous dye degradation. Among the tested samples, TiO2NRs0.3/CC achieved the most efficient photocatalytic activity, resulting in 90.2% degradation within 60 min. This superior performance highlights the crucial role of the TiO2–carbon interface, which enhances light absorption, promotes charge separation, and accelerates interfacial electron transfer. The normalized C/C0 plots in Figure 7b illustrate the degradation kinetics, with TiO2NRs0.3/CC consistently outperforming other composites. To quantify the reaction kinetics, the pseudo-first-order model was applied, and the corresponding ln(C0/C) curves are shown in Figure 7c. The calculated rate constants were 0.00216, 0.0122, 0.0177, 0.0332, and 0.0256 min−1 for CC, TiO2NRs, TiO2NRs0.1/CC, TiO2NRs0.3/CC, and TiO2NRs0.5/CC, respectively. The TiO2NRs0.3/CC sample exhibited the highest rate constant (0.0332 min−1), which is nearly three times greater than pristine TiO2NRs, demonstrating the strong influence of optimized CC loading on catalytic efficiency [30]. Such improvements have been consistently reported in TiO2–carbon composites, where the conductive support suppresses electron–hole recombination and provides additional adsorption sites for pollutant molecules [31]. Recyclability of the TiO2NRs0.3/CC photocatalyst is assessed across three consecutive cycles, as shown in Figure 7c. The MB degradation efficiency decreased slightly from 89.65% in the first run to 85.6% in the third cycle, indicating excellent structural stability and reusability of the composite electrode. The slight decline is attributed to surface fouling and partial active site deactivation during prolonged photocatalysis, in line with previous results [32]. For better context, Table 1 compares the TiO2NRs/CC photocatalysts with previously reported TiO2-based systems immobilized on carbon substrates. As seen in Table 1, the TiO2NR0.3/CC developed in this work demonstrates significantly higher H2 evolution rates and efficient dye degradation compared to earlier reports, underscoring the synergistic role of CC in enhancing photocatalytic performance.

2.7. Photocatalytic Mechanism

The enhanced photocatalytic activity of TiO2 NRs grown on CC is attributed to the synergistic interplay between the 1D semiconductor nanostructures and the conductive carbon substrate. Under simulated solar irradiation, rutile TiO2 NRs absorb photons with energies equal to or greater than their band gap, generating electron–hole (e/h+) pairs according to the following process:
T i O 2 + h v e C B + e V B +
The vertically aligned morphology of TiO2 NRs facilitates the rapid transport of photogenerated electrons along the c-axis toward the CC substrate. Due to its excellent conductivity, CC acts as an electron reservoir and mediator, effectively extracting and shuttling electrons while simultaneously suppressing charge recombination. This efficient separation and migration of charge carriers is confirmed by the observed PL quenching and reduced charge transfer resistance in EIS measurements.

2.7.1. H2 Evolution Pathway

The photogenerated electrons transferred from TiO2 to the CC are rapidly delivered to protons (H+) in the aqueous medium, leading to H2 evolution, as shown in the following equation.
2 H + + 2 e H 2
Simultaneously, the photogenerated holes are consumed by TEOA, working as a sacrificial electron donor, as shown by the following equation.
h + + T E A T E O A * + O x i d a t i o n   p r o d u c t s
Through this synergistic charge separation mechanism, the H2 generation efficiency is maximized, and TiO2NR0.3/CC achieves the highest rate (2.66 mmol h−1 g−1).

2.7.2. Dye Degradation Pathway

Under visible-light irradiation, TiO2 NRs generate charge carriers by following Equation 6. The electrons transferred to the CC reduce dissolved oxygen molecules to form reactive oxygen species (ROS), as shown in the following reactions,
O 2 + e O 2
O 2 + H + H O 2
2 H O 2 H 2 O 2 + O 2
H 2 O 2 + e O H + O H
Meanwhile, holes in TiO2 oxidize water molecules or hydroxyl ions to produce hydroxyl radicals, as shown below,
h + + H 2 O / O H   O H
The generated ROS ( O H , O 2 , H 2 O 2 ) and surface holes collectively degrade the methylene blue dye:
MB   +   ( O H / h + / O 2 )     C O 2 +   H 2 O + inorganic   ions
This combined oxidative and reductive pathway results in rapid MB removal. Especially, TiO2NR0.3/CC exhibits 90.2% degradation in 60 min.

3. Materials and Methodology

3.1. Materials

All chemicals were analytical grade from commercial suppliers. Tetrabutyl titanate (TBT, 98%, Alfa Aesar, Ward Hill, MA, USA) and titanium tetrachloride (TiCl4) were titanium precursors for TiO2 NRs synthesis. Carbon cloth from AvCarb, Lowell, MA, USA, was the substrate for TiO2 coating. Hydrochloric acid (HCl, 35–38%) was a pH adjuster and catalyst; ethanol (99.9%) and acetone (99.5%) were solvents and cleaning agents. Triethanolamine (TEOA, 99%) served as an electrolyte and sacrificial agent for H2 evolution and dye degradation. Methylene blue (MB, 95%) dye was used in degradation studies. All chemicals (i.e., TBT, TiCl4, HCl, ethanol, TEOS, acetone, TEOA, MB, RhB) were purchased from S D Fine Chem Limited, Mumbai, India.

3.2. Synthesis of TiO2NRs/CC

The TiO2 NRs were grown on CC by a hydrothermal method. Initially, the CC was refluxed in concentrated HNO3 acid at 120 °C for 4 h, during which the strong oxidizing conditions removed surface impurities and introduced oxygen-containing groups such as –COOH and –OH. These functional groups increased the hydrophilicity of the surface and provided active nucleation sites for TiO2 growth [39]. The cleaned and dried carbon fabric was then immersed overnight in ethanol solutions containing TiCl4 with concentrations of 0.1 M, 0.3 M, and 0.5 M. In ethanol, TiCl4 was partially hydrolyzed to form amorphous Ti–O–Ti species, which uniformly coated the fiber and acted as seed layers [40]. The seeded fabric was annealed in a muffle furnace at 400 °C for 30 min, transforming the amorphous titanium species into crystalline TiO2 nanocrystallites and improving adhesion to the carbon surface. In the next step, 4 mM of TBT, Ti(OC4H9)4 was dissolved in a mixed solvent of 20 mL hydrochloric acid and 20 mL acetone, and stirred to obtain a clear solution. Here, acetone served as a cosolvent to slow down uncontrolled hydrolysis, while HCl provided both water and chloride ions: water molecules initiated the controlled hydrolysis of TBT, and Cl ions selectively adsorbed onto crystal facets to promote anisotropic growth [22]. This precursor solution and the seed-coated carbon fabric were transferred to a Teflon-lined autoclave and heated at 180 °C for 6 h. These conditions are consistent with the literature reports showing that rutile TiO2 nanorods form preferentially in chloride-assisted acidic media under such conditions [41,42]. Under these circumstances, dissolved titanium species nucleated and preferentially grew into rod-like TiO2 structures on the seeded surface, guided by the high acidity and chloride concentrations [43]. After slow cooling to room temperature, the white TiO2 NRs were rinsed with DI water and ethanol to remove residual ions and organics, and then dried at 70 °C for 10 h to obtain a clean, uniform NRs array on the carbon fabric. The TiO2 NRs grown on carbon fabric were designated according to the concentration of the TiCl4 seed solution used during the seeding step. When 0.1 M TiCl4 was employed, the sample was denoted as TiONR0.1/CC, while those prepared with 0.3 M and 0.5 M TiCl4 seed solutions were labeled as TiONR0.3/CC and TiONR0.5/CC, respectively. This convention directly reflects the influence of the seeding concentration on the density and uniformity of TiO2 NR growth on the CC substrate. The CC was weighed before and after the deposition of TiO2 NRs. The bare CC was weighed before the growth of TiO2 NRs. After hydrothermal deposition and subsequent treatment, the TiO2 NRs NRs-decorated CC was dried and weighed again to determine the loading amount of TiO2. Pristine TiO2 NRs were synthesized without a CC substrate. The schematic illustration of the synthesis of TiO2 NRs on CC is presented in Figure 8, with the key reactions outlined below.
  • Hydrolysis of TiCl4 in ethanol (during seeding),
T i C l 4 + x H 2 O T i O x C l 4 x + x H C l
  • Hydrolysis of TBT (in precursor solution),
T i ( O C 4 H 9 ) 4 + 4 H 2 O T i ( O H ) 4 + 4 ( C 4 H 9 O H )
  • Condensation of hydroxide to TiO2,
T i ( O H ) 4 T i O 2 + 2 H 2 O

3.3. Characterization of TiO2NRs/CC

The crystalline structure of the photocatalyst powders was characterized using X-ray diffraction (XRD) with a Bruker D8 Advance X-ray diffractometer. X-ray spectra were produced using a sealed tube with filtered Cu Kα radiation (wavelength 1.54 Å) operated at 40 kV and 40 mA. The surface morphology, including shape and size, of the heterostructure nanomaterials was examined using field emission gun scanning electron microscopy (FEG-SEM, JEOL JSM-7600F, Tokyo, Japan) and high-resolution transmission electron microscopy (HR-TEM, Tecnai G2 F30, Minneapolis, MN, USA) at an accelerating voltage of 300 kV. Optical absorption properties of the photocatalysts were measured using a PerkinElmer Lambda 950 UV-Vis-NIR optical spectrophotometer (New York, NY, USA). The photoluminescence (PL) properties were evaluated using a Jasco V-770 spectrofluorometer (Tokyo, Japan) at an excitation wavelength of 390 nm under ambient conditions. Additionally, X-ray photoelectron spectroscopy (XPS) was performed using an AXIS Supra (Kratos Analytical, Stretford, UK) equipped with an Al Kα source to investigate the surface electronic states.

3.4. Photocatalytic Hydrogen (H2) Evolution Using TiO2NRs/CC

The photocatalytic H2 evolution activity of the as-prepared photocatalysts was evaluated in a 30 mL splitable borosilicate glass reactor under visible-light irradiation. A TiO2 NRs grown on a CC (3 cm diameter circle) was placed on a polyurethane foam ring to float on the surface of a 20 mL aqueous solution (90 vol% H2O, 10 vol% triethanolamine (TEOA) used as the electrolyte. The light source was a 400 W lamp (HPL-N 400 W) with a cutoff filter ≥420 nm. Before light irradiation, the reactor was purged with nitrogen gas for 30 min to remove dissolved oxygen and exposed to light irradiation with an intensity of 100 mW/cm2 measured by using an optical power meter (CEL-NP2000-2). The evolved H2 gas was quantified hourly using gas chromatography (Agilent GC 7820A) equipped with a thermal conductivity detector, with nitrogen (N2) as the carrier gas. No noble metal co-catalysts were added to enhance the photocatalytic H2 evolution rate. Additionally, the photocatalytic degradation of organic pollutants in water was investigated to assess the dual functional properties of the photocatalytic substrate. The photodegradation of methylene blue (MB) dye was conducted under the same light source and irradiation intensity of the same light source, using the same TiO2 NRs grown on CC for H2 evolution. The photocatalytic substrate was floated on 100 mL of 0.01 gL−1 MB dye solution. Before light exposure, the dye solution was stirred magnetically in dark conditions for 20 min to establish adsorption–desorption equilibrium. After specific intervals of light irradiation, the reaction solution was analyzed using a UV–Vis spectrophotometer (centered at a wavelength of 664 nm). The photocatalytic dye degradation efficiency (η) was calculated using the following equation:
η = C 0 C t C 0 × 100 %
where C 0 is the initial concentration of MB dye after adsorption–desorption equilibrium, and C t is the concentration of dye after a certain time under light irradiation.

3.5. Electrochemical Measurement

Electrochemical experiments were conducted using a PGSTAT 302N workstation configured with a standard three-electrode system, comprising an Ag/AgCl (3 M KCl) reference electrode for stable potential, a platinum wire counter electrode to complete the circuit, and a TiO2NRs/CC working electrode. The TiO2NRs/CC was submerged in a 0.5 M Na2SO4 aqueous electrolyte solution (pH 6.8) to maintain consistent experimental conditions. Electrochemical impedance spectroscopy (EIS) was performed over a frequency range of 100 kHz to 10 mHz to evaluate charge transfer and impedance characteristics. Mott–Schottky analysis was conducted to measure the flat-band potential and charge carrier density of the TiO2NRs/CC electrode, critical for understanding its semiconductor properties. This analysis was executed at a fixed frequency of 1 kHz.

4. Conclusions

In this study, rutile TiO2 NRs arrays were successfully synthesized for the first time on CC substrates via a TiCl4-assisted hydrothermal route, with different seeding concentrations. The structural and spectroscopic analyses confirmed the formation of vertically aligned, single-crystalline rutile TiO2 nanorods strongly anchored to the conductive carbon framework. Among the synthesized samples, TiO2NR0.3/CC exhibited the most favorable balance between nanorod density, interfacial contact, and charge transport characteristics. Studies on optical studies revealed a red-shifted absorption edge and narrowed band gap relative to bare TiO2 NRs; in addition, photoluminescence quenching and electrochemical impedance spectroscopy confirmed the efficient suppression of electron–hole recombination and reduced interfacial resistance. These enhancements give rise to superior photocatalytic activity, achieving a maximum H2 evolution and rate with TiO2NR0.3/CC photocatalysts, along with outstanding cyclic stability across multiple cyclic operations. Parallel degradation studies on TiO2NR0.3/CC of methylene blue under visible light demonstrate 90.2% dye removal within 60 min, while maintaining good activity after three consecutive cycles. These results unequivocally demonstrate that CC serves as a passive support and as an active electron mediator, enhancing charge separation and promoting durable photocatalysis. The synergistic integration of TiO2 NRs with CC effectively addresses the intrinsic drawbacks of pristine TiO2 NRs, including limited solar absorption, poor charge transport, and difficulties in catalyst recovery. The integration of 1D TiO2 NRs with flexible CC would provide a recyclable, structurally robust, and high-performance photocatalytic platform. Thus, the present work establishes a new promising pathway for the design of multifunctional photocatalysts for solar-driven H2 production and wastewater remediation, with broad implications for energy conversion and environmental sustainability.

Author Contributions

Conceptualization, methodology, data curation, and writing—original, S.R.A.; formal analysis, validation, funding acquisition, writing—review and editing, K.B.A.; formal analysis, writing review and editing, S.J.L.; formal analysis, writing—review and editing, N.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through a Large Research Project under grant number (RGP. 2/236/46). Also, we give thanks to the National Research Foundation of Korea (NRF) grant funded by the Korea government (MIST) (RS-2024-00341278).

Data Availability Statement

The data will be made available by the authors on request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  2. Shakeelur Raheman, A.R.; Mane, R.S.; Wilson, H.M.; Jha, N. CdSe quantum dot/white graphene hexagonal porous boron nitride sheet (h-PBNs) heterostructure photocatalyst for solar driven H2 production. J. Mater. Chem. C Mater. 2021, 9, 8524–8536. [Google Scholar] [CrossRef]
  3. Joshy, D.; Narendranath, S.B.; Ismail, Y.A.; Periyat, P. Recent progress in one dimensional TiO2 nanomaterials as photoanodes in dye-sensitized solar cells. Nanoscale Adv. 2022, 4, 5202–5232. [Google Scholar] [CrossRef] [PubMed]
  4. Ghibaudo, N.; Ferretti, M.; Al-Hetlani, E.; Madkour, M.; Amin, M.O.; Alberti, S. Synthesis and characterization of TiO2-based supported materials for industrial application and recovery in a pilot photocatalytic plant using chemometric approach. Environ. Sci. Pollut. Res. 2024, 31, 20556–20567. [Google Scholar] [CrossRef]
  5. Fernández-Ibáñez, P.; Blanco, J.; Malato, S.; De Las Nieves, F.J. Application of the colloidal stability of TiO2 particles for recovery and reuse in solar photocatalysis. Water Res. 2003, 37, 3180–3188. [Google Scholar] [CrossRef]
  6. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B 2012, 125, 331–349. [Google Scholar] [CrossRef]
  7. Murthy, R.; Neelakantan, S.C. Graphitic Carbon Cloth-Based Hybrid Molecular Catalyst: A Non-conventional, Synthetic Strategy of the Drop Casting Method for a Stable and Bifunctional Electrocatalyst for Enhanced Hydrogen and Oxygen Evolution Reactions. ACS Omega 2022, 7, 32604–32614. [Google Scholar] [CrossRef]
  8. Xi, X.; Lu, Z.; Chen, H.J.; Tian, W.; Yang, Y.L. Flexible carbon-fiber/g-C3N4 nanosheet array as recyclable photocatalysts and photoelectrocatalysts. Sci. Rep. 2025, 15, 21090. [Google Scholar] [CrossRef]
  9. Juang, S.E.; Chin, N.C.; Chang, Y.C.; Chou, C.M. Fabrication of ZnCo2O4-Zn(OH)2 Microspheres on Carbon Cloth for Photocatalytic Decomposition of Tetracycline. Molecules 2024, 29, 4054. [Google Scholar] [CrossRef] [PubMed]
  10. Chang, C.J.; Kao, Y.C.; Lin, K.S.; Chen, C.Y.; Kang, C.W.; Yang, T.H. Carbon fiber cloth@BiOBr/CuO as immobilized membrane-shaped photocatalysts with enhanced photocatalytic H2 production activity. J. Taiwan. Inst. Chem. Eng. 2023, 149, 104998. [Google Scholar] [CrossRef]
  11. Wan, W.; Li, Y.; Bai, S.; Yang, X.; Chi, M.; Shi, Y.; Liu, C.; Zhang, P. Three-Dimensional Porous ZnO-Supported Carbon Fiber Aerogel with Synergistic Effects of Adsorption and Photocatalysis for Organics Removal. Sustainability 2023, 15, 13088. [Google Scholar] [CrossRef]
  12. Jiang, L.; Gao, X.; Chen, S.; Ashok, J.; Kawi, S. Oxygen-Deficient WO3/TiO2/CC Nanorod Arrays for Visible-Light Photocatalytic Degradation of Methylene Blue. Catalysts 2021, 11, 1349. [Google Scholar] [CrossRef]
  13. Yi, Y.; Guan, Q.; Wang, W.; Jian, S.; Li, H.; Wu, L.; Zhang, H.; Jiang, C. Recyclable Carbon Cloth-Supported ZnO@Ag3PO4 Core–Shell Structure for Photocatalytic Degradation of Organic Dye. Toxics 2023, 11, 70. [Google Scholar] [CrossRef]
  14. Jian, S.; Wang, W.; Wu, L.; Li, H.; Hong, C.; Long, S.; Zhou, W.; Guo, Y. Construction of highly efficient carbon cloth-supported S-scheme Co3O4/AgIO4 heterojunction for photocatalytic degradation of Rhodamine B organic dye. J. Photochem. Photobiol. A Chem. 2024, 452, 115598. Available online: https://ssrn.com/abstract=4729376 (accessed on 4 September 2025). [CrossRef]
  15. He, D.; Li, X.; Jia, B.; Li, M.; Tang, H.; Li, H.; Ma, Y.; Wang, X. Photocatalytic removal of volatile organic compounds by monolithic heterojunction materials: A review. Catal. Sci. Technol. 2025, 15, 5202–5225. [Google Scholar] [CrossRef]
  16. Shi, H.; Wen, G.; Nie, Y.; Zhang, G.; Duan, H. Flexible 3D carbon cloth as a high-performing electrode for energy storage and conversion. Nanoscale 2020, 12, 5261–5285. [Google Scholar] [CrossRef]
  17. Ferrari, A.; Robertson, J. Interpretation of Raman spectra of disordered and amorphous carbon. Phys. Rev. B 2000, 61, 14095. [Google Scholar] [CrossRef]
  18. Hanaor, D.A.H.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2010, 46, 855–874. [Google Scholar] [CrossRef]
  19. Yang, H.G.; Zeng, H.C. Preparation of hollow anatase TiO2 nanospheres via Ostwald ripening. J. Phys. Chem. B 2004, 108, 3492–3495. [Google Scholar] [CrossRef]
  20. Chen, C.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  21. Kavan, L.; Grätzel, M.; Gilbert, S.E.; Klemenz, C.; Scheel, H.J. Electrochemical and photoelectrochemical investigation of single-crystal anatase. J. Am. Chem. Soc. 1996, 118, 6716–6723. [Google Scholar] [CrossRef]
  22. Liu, B.; Aydil, E.S. Growth of oriented single-crystalline rutile TiO2 nanorods on transparent conducting substrates for dye-sensitized solar cells. J. Am. Chem. Soc. 2009, 131, 3985–3990. [Google Scholar] [CrossRef]
  23. Bavykin, D.V.; Parmon, V.N.; Lapkin, A.A.; Walsh, F.C. The effect of hydrothermal conditions on the mesoporous structure of TiO2 nanotubes. J. Mater. Chem. 2004, 14, 3370–3377. [Google Scholar] [CrossRef]
  24. Ji, Y. Facile route for synthesis of TiO2 nanorod arrays by high-temperature calcinations. Mater. Lett. 2013, 108, 208–211. [Google Scholar] [CrossRef]
  25. Yu, J.; Yu, H.; Cheng, B.; Zhou, M.; Zhao, X. Enhanced photocatalytic activity of TiO2 powder (P25) by hydrothermal treatment. J. Mol. Catal. A Chem. 2006, 253, 112–118. [Google Scholar] [CrossRef]
  26. Pakpahan, S.; Gultom, R. Optimizing Photovoltaic Performance in p-Cu2O/n-TiO2 Heterojunction Solar Cells: The Impact of Annealing Temperature, Layer Thickness, and Carbon Doping. Indones. J. Chem. Res. 2025, 13, 23–29. [Google Scholar] [CrossRef]
  27. Negoescu, D.; Bratan, V.; Gherendi, M.; Atkinson, I.; Culita, D.C.; Neacsu, A.; Baran, A.; Petrescu, S.; Parvulescu, V. Iron Promoted TiO2-Activated Carbon Nanocomposites for Photocatalytic Degradation of Congo Red in Water. Catalysts 2024, 14, 844. [Google Scholar] [CrossRef]
  28. Lee, S.F.; Jimenez-Relinque, E.; Martinez, I.; Castellote, M. Effects of Mott–Schottky Frequency Selection and Other Controlling Factors on Flat-Band Potential and Band-Edge Position Determination of TiO2. Catalysts 2023, 13, 1000. [Google Scholar] [CrossRef]
  29. Trinh, N.B.; Nguyen, T.A.; Van Vu, S.; Vo, H.G.T.; Lo, T.N.H.; Park, I.; Vo, K.Q. Modified hydrothermal method for synthesizing titanium dioxide-decorated multiwalled carbon nanotube nanocomposites for the solar-driven photocatalytic degradation of dyes. RSC Adv. 2024, 14, 34037–34050. [Google Scholar] [CrossRef]
  30. Sriramoju, J.B.; Muniyappa, M.; Marilingaiah, N.R.; Sabbanahalli, C.; Shetty, M.; Mudike, R.; Chitrabanu, C.P.; Shivaramu, P.D.; Nagaraju, G.; Rangappa, K.S.; et al. Carbon-based TiO2-x heterostructure nanocomposites for enhanced photocatalytic degradation of dye molecules. Ceram. Int. 2021, 47, 10314–10321. [Google Scholar] [CrossRef]
  31. Hossen, M.A.; Ikreedeegh, R.R.; Aziz, A.A.; Zerga, A.Y.; Tahir, M. Carbon-based nanomaterials (CNMs) modified TiO2 nanotubes (TNTs) photo-driven catalysts for sustainable energy and environmental applications: A comprehensive review. J. Env. Chem. Eng. 2024, 12, 114088. [Google Scholar] [CrossRef]
  32. Zhang, Z.; Wang, C.-C.; Zakaria, R.; Ying, J.Y. Role of Particle Size in Nanocrystalline TiO2-Based Photocatalysts. J. Phys. Chem. B 1998, 102, 10871–10878. [Google Scholar] [CrossRef]
  33. Xia, W.; Qian, H.; Zeng, X.; Sun, J.; Wang, P.; Luo, M.; Dong, J. TiO2@Sn3O4 nanorods vertically aligned on carbon fiber papers for enhanced photoelectrochemical performance. RSC Adv. 2019, 9, 23334–23342. [Google Scholar] [CrossRef]
  34. Tahir, M.; Alesayi, M.T.H.; Alshehhi, S.M.S. Recycling Carbon Fiber-Reinforced Polymers (CFRPs) to Construct CFs/TiO2 Nanotexture with Efficient Interface Charge Transfer for Stimulating Photocatalytic Hydrogen Production. Energy Fuels 2023, 37, 12319–12334. [Google Scholar] [CrossRef]
  35. Behpour, M.; Shirazi, P.; Rahbar, M. Immobilization of the Fe2O3/TiO2 photocatalyst on carbon fiber cloth for the degradation of a textile dye under visible light irradiation, Reaction Kinetics. Mech. Catal. 2019, 127, 1073–1085. [Google Scholar] [CrossRef]
  36. Wang, Y.; Omsinsombon, J.; Prateepmaneerak, N.; Li, S.; Zhu, G.; Khan, M.Z.; Taghavian, H.; Venkataraman, M.; Militký, J.; Chaiyasat, A. UV-activated TiO2-coated carbon felt for photocatalytic dye degradation and antibacterial applications. Surf. Interfaces 2025, 60, 106018. [Google Scholar] [CrossRef]
  37. Chennam, P.K.; Sepúlveda, M.; Rihova, M.; Alijani, M.; Kachlík, M.; Zazpe, R.; Pavlinak, D.; Maca, K.; Macak, J.M. Carbon fibers decorated with TiO2 nanoparticles for photocatalytic degradation of methylene blue dye. Front. Nanotechnol. 2024, 6, 1483917. [Google Scholar] [CrossRef]
  38. Tian, Q.; Yi, S.; Li, C.; Liu, Y.; Niu, Z.; Yue, X.; Liu, Z. Design of charge transfer channels: Defective TiO2/MoP supported on carbon cloth for solar-light-driven hydrogen generation. Inorg. Chem. Front. 2021, 8, 2017–2026. [Google Scholar] [CrossRef]
  39. Wang, L.; Liu, N.; Guo, Z.; Wu, D.; Chen, W.; Chang, Z.; Yuan, Q.; Hui, M.; Wang, J. Nitric Acid-Treated Carbon Fibers with Enhanced Hydrophilicity for Candida tropicalis Immobilization in Xylitol Fermentation. Materials 2016, 9, 206. [Google Scholar] [CrossRef]
  40. Tamilselvan, V.; Yuvaraj, D.; Kumar, R.R.; Rao, K.N. Growth of rutile TiO2 nanorods on TiO2 seed layer deposited by electron beam evaporation. Appl. Surf. Sci. 2012, 258, 4283–4287. [Google Scholar] [CrossRef]
  41. Guo, W.; Xu, C.; Wang, X.; Wang, S.; Pan, C.; Lin, C.; Wang, Z.L. Rectangular Bunched Rutile TiO2 Nanorod Arrays Grown on Carbon Fiber for Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2012, 134, 4437–4441. [Google Scholar] [CrossRef] [PubMed]
  42. Diao, W.; He, J.; Wang, Q.; Rao, X.; Zhang, Y. Na and Cl co-doped TiO2 nanorod arrays on carbon cloth for efficient photocatalytic degradation of formaldehyde under UV/visible LED irradiation. Catal. Sci. Technol. 2021, 11, 230–238. [Google Scholar] [CrossRef]
  43. Han, X.; Kuang, Q.; Jin, M.; Xie, Z.; Zheng, L. Synthesis of titania nanosheets with a high percentage of exposed (001) facets and related photocatalytic properties. J. Am. Chem. Soc. 2009, 131, 3152–3153. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD pattern of CC and TiO2 NRs grown on CC.
Figure 1. XRD pattern of CC and TiO2 NRs grown on CC.
Catalysts 15 00961 g001
Figure 2. (a,b) SEM images of pristine CC showing smooth fiber morphology, (c) digital photograph of TiO2NR0.3/CC, (df) SEM images of TiO2NR0.3/CC at different magnifications, (g) TEM image of TiO2 NRs, (h) high-resolution TEM image of TiO2 NRs, (i) shows the SAED pattern.
Figure 2. (a,b) SEM images of pristine CC showing smooth fiber morphology, (c) digital photograph of TiO2NR0.3/CC, (df) SEM images of TiO2NR0.3/CC at different magnifications, (g) TEM image of TiO2 NRs, (h) high-resolution TEM image of TiO2 NRs, (i) shows the SAED pattern.
Catalysts 15 00961 g002
Figure 3. (a) Full survey XPS spectra of CC and TiO2NRs0.3/CC, (bd) high-resolution spectra of Ti 2p, O 1s, and C1s.
Figure 3. (a) Full survey XPS spectra of CC and TiO2NRs0.3/CC, (bd) high-resolution spectra of Ti 2p, O 1s, and C1s.
Catalysts 15 00961 g003
Figure 4. (a,b) UV–Vis DRS spectra of photocatalysts, (c) Tauc plot, and (df) deconvoluted photoluminescence spectra of photocatalyst.
Figure 4. (a,b) UV–Vis DRS spectra of photocatalysts, (c) Tauc plot, and (df) deconvoluted photoluminescence spectra of photocatalyst.
Catalysts 15 00961 g004
Figure 5. (a) EIS spectra, (b,c) Mott–Schottky plots of TiO2 NRs and TiO2NRs0.3M/CC, (d) linear sweep voltammetry measurement of electros.
Figure 5. (a) EIS spectra, (b,c) Mott–Schottky plots of TiO2 NRs and TiO2NRs0.3M/CC, (d) linear sweep voltammetry measurement of electros.
Catalysts 15 00961 g005
Figure 6. (a) Time course H2 evolution of photocatalysts, (b) H2 production rate, (c) photocatalytic cyclic stability of TiO2NRs0.3/CC.
Figure 6. (a) Time course H2 evolution of photocatalysts, (b) H2 production rate, (c) photocatalytic cyclic stability of TiO2NRs0.3/CC.
Catalysts 15 00961 g006
Figure 7. (a) Time-dependent absorption spectra of MB dye using TiO2NRs0.3/CC, (b) photocatalytic degradation curve of as-prepared photocatalysts, (c) kinetics of MB dye degradation, (d) recycling stability of TiO2NRs0.3/CC photocatalyst.
Figure 7. (a) Time-dependent absorption spectra of MB dye using TiO2NRs0.3/CC, (b) photocatalytic degradation curve of as-prepared photocatalysts, (c) kinetics of MB dye degradation, (d) recycling stability of TiO2NRs0.3/CC photocatalyst.
Catalysts 15 00961 g007
Figure 8. Schematic illustration of the synthesis of TiO2NRs on CC substrate.
Figure 8. Schematic illustration of the synthesis of TiO2NRs on CC substrate.
Catalysts 15 00961 g008
Table 1. TiO2 NRs on CC or carbon fiber paper (CFP), for photocatalytic H2 evolution and photocatalytic dye degradation.
Table 1. TiO2 NRs on CC or carbon fiber paper (CFP), for photocatalytic H2 evolution and photocatalytic dye degradation.
Photocatalyst
System
Light
Source
Photocatalytic H2 EvolutionPhotocatalytic Dye
Degradation
Reference
TiO2@Sn3O4 nanorodsVisible-light PEC5.23 µmol h−1 (vs. 1.13 µmol h−1 for TiO2 NRs) [33]
Carbon fiber-reinforced polymers/TiO2 composite (powdered CFs mixed with TiO2)UV–Vis, MeOH sacrificial85.4 µmol g−1 h−1 (3 wt% CFs/TiO2), 2.87× vs. bare TiO2 [34]
Fe2O3/TiO2 immobilized on CFCVisible light 97.54% color removal; ~84% COD removal (240 min)[35]
UV-activated TiO2 coated felt (covalently anchored)UV 95% Orange II in 180 min[36]
TiO2 nanoparticles decorated on CFsUV (λ ≈ 365 nm) First-order kinetics[37]
Defective TiO2/MoP heterojunction on CCSolar light (panel; immobilized sheet)14.7 µmol h−1 cm−2; ~19× higher than TiO2/CC [38]
TiO2 nanorods (powder)Vis (λ ≥ 400 nm), TEOA0.66 mmol h−1 g−1 Present research
TiO2NR0.1/CCVis (λ ≥ 400 nm), TEOA0.90 mmol h−1 g−1 Present research
TiO2NR0.3/CCVis (λ ≥ 400 nm), TEOA2.66 mmol h−1 g−1TiO2NR0.3/CC achieved 90.2% in 90 minPresent research
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

AR, S.R.; Ansari, K.B.; Lee, S.J.; Salunke, N. Synergistic Integration of TiO2 Nanorods with Carbon Cloth for Enhanced Photocatalytic Hydrogen Evolution and Wastewater Remediation. Catalysts 2025, 15, 961. https://doi.org/10.3390/catal15100961

AMA Style

AR SR, Ansari KB, Lee SJ, Salunke N. Synergistic Integration of TiO2 Nanorods with Carbon Cloth for Enhanced Photocatalytic Hydrogen Evolution and Wastewater Remediation. Catalysts. 2025; 15(10):961. https://doi.org/10.3390/catal15100961

Chicago/Turabian Style

AR, Shakeelur Raheman, Khursheed B. Ansari, Sang Joon Lee, and Nilesh Salunke. 2025. "Synergistic Integration of TiO2 Nanorods with Carbon Cloth for Enhanced Photocatalytic Hydrogen Evolution and Wastewater Remediation" Catalysts 15, no. 10: 961. https://doi.org/10.3390/catal15100961

APA Style

AR, S. R., Ansari, K. B., Lee, S. J., & Salunke, N. (2025). Synergistic Integration of TiO2 Nanorods with Carbon Cloth for Enhanced Photocatalytic Hydrogen Evolution and Wastewater Remediation. Catalysts, 15(10), 961. https://doi.org/10.3390/catal15100961

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop