Next Article in Journal
Photocatalytic Degradation of Ant Trail Pheromones by P25 TiO2
Previous Article in Journal
Enhanced Catalytic Performance of PdOx/CuO Derived from Pd-Embedded CuBTC for Oxidative Carbonylation of Phenol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interfacial Engineering of CN-B/Ti3C2 MXene Heterojunction for Synergistic Solar-Driven CO2 Reduction

1
School of Chemistry and Chemical Engineering, Institute for Advanced Materials, Jingjiang College, Jiangsu University, Zhenjiang 212013, China
2
Department of Chemistry, Faculty of Science and Data Analytics, Sepuluh Nopember Institute of Technology, Surabaya 60111, Indonesia
3
Department of Biology, Hong Kong Baptist University, Hong Kong SAR, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1037; https://doi.org/10.3390/catal15111037 (registering DOI)
Submission received: 18 September 2025 / Revised: 18 October 2025 / Accepted: 28 October 2025 / Published: 2 November 2025
(This article belongs to the Special Issue Recent Advances in Photo/Electrocatalytic CO2 Reduction)

Abstract

Photocatalytic CO2 reduction holds great potential for sustainable solar fuel production, yet its practical application is often limited by inefficient charge separation and poor product selectivity. The photothermal effect presents a viable strategy to address these challenges by reducing activation energies and accelerating reaction kinetics. In this work, we report a rationally designed CN-B/Ti3C2 heterojunction that effectively leverages photothermal promotion for enhanced CO2 reduction. The black carbon nitride (CN-B) framework, synthesized via a one-step calcination of urea and Phloxine B, exhibits outstanding photothermal conversion, reaching 131.4 °C under 300 mW cm−2 illumination, which facilitates CO2 adsorption and charge separation. Coupled with Ti3C2 MXene, the optimized composite (3:1) achieves remarkable CO and CH4 production rates of 80.21 and 35.13 μmol g−1 h−1, respectively, without any cocatalyst—representing a 2.9-fold and 8.8-fold enhancement over CN-B and g-C3N4 in CO yield. Mechanistic studies reveal that the improved performance stems from synergistic effects: a built-in electric field prolongs charge carrier lifetime (3.15 ns) and reduces interfacial resistance, while localized heating under full-spectrum light further promotes CO2 activation. In situ Fourier transform infrared (FTIR) spectroscopy confirms the accelerated formation of key intermediates (*COOH and *CO). The catalyst also maintains excellent stability over 24 h. This study demonstrates the promise of combining photothermal effects with heterojunction engineering for efficient and durable CO2 photoreduction.

1. Introduction

The escalating atmospheric CO2 concentration has precipitated a global climate crisis, necessitating urgent innovations in solar-driven carbon capture and conversion technologies [1,2,3]. Photocatalytic CO2 reduction represents a promising route to mitigate emissions while advancing renewable energy economies, yet conventional semiconductors face intrinsic limitations in solar spectrum utilization (limited visible-to-infrared photon utilization efficiency), rapid charge recombination, and sluggish surface reaction kinetics, restricting their practical deployment [4,5,6]. Graphitic carbon nitride (g-C3N4), a visible-light-active semiconductor with a tunable bandgap (~2.7 eV), has garnered significant attention due to its low cost, non-toxicity, and robustness. Nevertheless, pristine g-C3N4 suffers from inherent limitations, including a narrow light absorption edge (~450 nm), high exciton binding energy, and weak CO2 adsorption, leading to low quantum efficiency and poor product selectivity [7,8,9]. While established strategies like elemental doping and morphology engineering have significantly improved its performance, many have not fully explored the utilization of infrared photons (>700 nm), which constitute nearly half of the solar spectrum. Additionally, the deliberate integration of photothermal effects remains a relatively underexplored avenue for further enhancing catalytic efficiency through reduced activation barriers [10,11,12].
Recent advances in photothermal-assisted photocatalysis have revolutionized this field by synergizing light absorption and thermal activation [13,14]. Lin et al. developed a sulfur-vulcanized Ti3C2 MXene that achieves efficient CO2-to-C2+ fuel conversion (C2H4: 3.55 mmol g−1 h−1) via full-spectrum photothermal-photocatalysis, setting a new benchmark with 0.045% solar-to-fuel efficiency [15]. This approach leverages materials’ ability to convert underutilized infrared photons into localized heat, which weakens the C=O bond (≈750 kJ mol−1) and accelerates electron injection into CO2 antibonding orbitals, thereby enabling kinetically hindered multi-electron reduction pathways [16]. Concurrently, two-dimensional transition metal carbides (MXenes), exemplified by Ti3C2, have emerged as game-changing cocatalysts [17,18,19]. Ti3C2 MXene exhibits metallic conductivity, broad-spectrum light absorption (UV to NIR), and exceptional photothermal conversion efficiency. Its hydrophilic surface terminations (-O, -OH) enhance CO2 adsorption capacity, while layered structures provide abundant redox-active sites [20,21]. When integrated into Schottky junctions, Ti3C2’s higher work function creates built-in electric fields that drive directional electron transfer, substantially reducing recombination losses [22,23,24,25]. Despite these merits, existing MXene-modified g-C3N4 systems lack infrared-specific sensitization, limiting full-spectrum utilization, and fail to unify thermal and electronic optimization in a single architecture.
Herein, we engineer a novel CN-B/Ti3C2 Schottky heterojunction to synergize broadband light harvesting, photothermal catalysis, and interfacial charge dynamics. The CN-B matrix is synthesized via the one-step calcination of g-C3N4 with phloxine B dye—a red organic sensitizer that induces partial heptazine ring opening, extending visible-light absorption to 800 nm and amplifying photothermal conversion (ΔT > 35 °C under 1-sun illumination) [26]. Ti3C2 MXene is integrated via hydrothermal assembly, serving dual roles: as a Schottky booster facilitating electron-hole separation via built-in electric fields, and as a photothermal amplifier converting infrared photons into localized heat to thermally excite CO2 molecules and lower activation barriers for multi-electron reduction [27,28,29]. This dual functionality creates a dynamic reaction microenvironment where thermal energy synergizes with optimized charge separation to promote •CO2 intermediate formation, a critical step for selective CH4 generation [30,31].
The resultant heterojunction achieves excellent CO2 reduction performance, CO and CH4 yields of 80.21 μmol g−1 h−1 and 35.13 μmol g−1 h−1, respectively—representing 2.9-fold and 3.2-fold enhancements over pristine CN-B. In situ DRIFTS data confirm thermally accelerated C=O dissociation and stabilization of •CO2 intermediates, while the system maintains >90% activity over 24 h due to the prevention of oxidative corrosion by hydrophobic MXene layers. This work pioneers three innovations: phloxine B sensitization for full-spectrum infrared harvesting, Ti3C2-enabled thermo-electronic synergy, and waste-free design avoiding noble metals. By unifying photothermal heating, carrier dynamics, and targeted CO2 activation, this study establishes a scalable blueprint for solar-powered carbon neutrality technologies.

2. Results and Discussion

2.1. Photocatalyst Characterization

Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) analyses reveal the distinct morphologies and interfacial characteristics of the synthesized materials. As shown in Figure 1a, CN-B exhibits a porous, layered stacking structure, where the open pore architecture facilitates active site exposure. Multilayered Ti3C2 nanosheets, obtained via HF etching of Ti3AlC2, display a characteristic two-dimensional layered morphology (Figure 1b). The SEM image of the CN-B/Ti3C2 (3:1) heterostructure (Figure 1c) demonstrates CN-B intimately attached to the layered surface of Ti3C2, forming a well-integrated interface. This unique microstructure leverages the inherent advantages of MXene materials—high specific surface area and excellent conductivity—while the heterojunction formation promotes efficient separation and transport of photogenerated charge carriers, providing an ideal platform for enhanced photocatalytic activity [31,32].
TEM analysis further elucidates the fine structural details. Figure 1e clearly shows the typical layered structure of Ti3C2 with ordered stacking. Notably, the CN-B/Ti3C2 (3:1) composite (Figure 1f) retains the characteristic curved, porous nanosheet morphology of CN-B (Figure 1d) while achieving effective integration of both phases. Elemental mapping (Figure 1g–j) confirms the spatial distribution of characteristic elements: C and N signals correspond precisely to the CN-B nanosheet structure, while Ti signals localize exclusively to the Ti3C2 regions. The elements exhibit a uniform, interpenetrating 3D spatial distribution, with overlapping C/N and Ti signal clouds forming a continuous, non-segregated coverage across the composite surface. This indicates intimate atomic-scale interfacial coupling between CN-B and Ti3C2 nanosheets. Such uniform dispersion prevents agglomeration of individual components and maximizes heterointerface contact area, thus enabling efficient cross-phase charge carrier migration. From a materials chemistry perspective, this microscale elemental homogeneity provides direct structural evidence for stable heterojunction interface formation, underpinning subsequent analyses of band alignment, charge transfer kinetics, and synergistic catalytic site interactions [33,34].

2.2. Surface Composition and Photoelectric Analysis

X-ray diffraction (XRD) was employed to characterize the crystallographic phases of the photocatalysts. Figure 2a displays XRD patterns of CN-B, Ti3C2, and CN-B/Ti3C2 composites with varying mass ratios. CN-B exhibits two distinct diffraction peaks at 13.1° and 27.2°, corresponding to the (001) plane (reflecting in-plane triazine ring ordering) and interlayer stacking along the c-axis, respectively [35]. Multilayered Ti3C2 obtained via HF etching shows characteristic peaks at 18.28°, 25.8°, 36°, and 59.2°, indexed to the (004), (006), (008), and (110) planes. The composite patterns confirm successful hybridization, retaining CN-B peaks at 13.1°/27.2° and Ti3C2 peaks at 18.28–59.2°. Notably, increasing the CN-B/Ti3C2 ratio progressively narrows the full-width at half-maximum (FWHM) of the dominant diffraction peaks, correlating with enhanced crystallinity. The optimal CN-B/Ti3C2 (3:1) composite demonstrates the sharpest peaks, indicating superior crystallinity [36]. Fourier-transform infrared (FTIR) spectra (Figure 2b) reveal consistent chemical bonding across CN-B and all composites. Characteristic absorption bands include: 810 cm−1 (out-of-plane breathing vibration of triazine units), 1200–1700 cm−1 (aromatic C-N stretching modes of heterocycles), and 3000–3500 cm−1 (N-H stretching vibrations). The spectral fidelity confirms structural integrity of CN-B within the composites [37].
X-ray photoelectron spectroscopy (XPS) was employed to probe the elemental composition and chemical states of CN-B and the CN-B/Ti3C2 (3:1) composite. The survey spectrum (Figure 3a) confirms the co-presence of C, N, and Ti in the composite, indicating successful hybridization. In the high-resolution N 1s spectrum (Figure 3c), the peaks are deconvoluted into three components at 395.65 eV, 397.52 eV, and 401.60 eV, corresponding to sp2 C-N═C, sp3 N-(C)3, and C-N-H species, respectively. A positive binding energy shift in N 1s relative to pristine CN-B suggests electron transfer from CN-B to Ti3C2, which enhances charge separation efficiency. Similarly, the C 1s spectrum (Figure 3b) of the composite exhibits peaks at 284.8 eV (C-C), 286.1 eV (C-N-H), and 288.1 eV (N-C═N), along with additional features at 281.7 eV, attributed to C-Ti bonding, and at 288.6 eV, assigned to C-O species. These also display a noticeable shift toward higher binding energies, further supporting electron migration toward Ti3C2. The Ti 2p spectrum (Figure 3d) shows doublet peaks at 455.3 eV (Ti-C 2p3/2), 458.7 eV (Ti-O 2p3/2), 461.1 eV (Ti-C 2p1/2), and 464.2 eV (Ti-O 2p1/2). A consistent redshift in these Ti 2p peaks, compared to pure Ti3C2, reflects an altered electronic environment and strong interfacial coupling. This synergy establishes Ti3C2 as an effective electron acceptor, facilitating charge transfer across the heterojunction. The built-in electric field induced by interfacial band alignment significantly suppresses electron-hole recombination, thereby optimizing the photocatalytic CO2 reduction performance [38].
The light absorption properties of the photocatalysts were investigated by UV-vis diffuse reflectance spectroscopy (DRS). As shown in Figure 4a, both CN-B and CN-B/Ti3C2 exhibit broad absorption across the visible to near-infrared range (450–800 nm). Incorporation of Ti3C2 induces a notable red-shift in the absorption edge, indicating enhanced light-harvesting capability. The absorption intensity increases systematically with higher CN-B content, reaching a maximum at the optimal CN-B/Ti3C2 mass ratio of 3:1. The bandgap energies were determined from Tauc plots (Figure 4b) using the equation (αhυ)2 = k(hυ − Eg). CN-B and CN-B/Ti3C2 (3:1) exhibit direct bandgaps of 2.88 eV and 2.82 eV, respectively. The reduced bandgap of the composite facilitates greater visible-light-driven electron generation. XPS valence band spectra further reveal a shift in the valence band maximum from 1.55 eV (CN-B) to 1.46 eV (CN-B/Ti3C2), indicating enhanced reducibility. Using the equation EVB = ECB + Eg, the calculated conduction band positions are −1.33 eV for CN-B and −1.36 eV for CN-B/Ti3C2. The more negative conduction band potential of the composite demonstrates superior electron reduction capacity, consistent with its improved photocatalytic performance.
Nitrogen adsorption–desorption isotherms (Figure 5a) reveal type-IV hysteresis loops in the P/P0 range of 0.5–1.0 for CN-B, Ti3C2, and the CN-B/Ti3C2 (3:1) composite, confirming their mesoporous structures. Pore size distribution curves further exhibit prominent peaks within 1–20 nm, consistent with mesoporous characteristics. The interconnected porous network significantly increases active site accessibility, thereby enhancing catalytic reaction kinetics. The composite material demonstrates markedly optimized surface properties, exhibiting a specific surface area of 60.454 m2 g−1—nearly double that of pristine CN-B (31.39 m2 g−1). This expansion in surface area not only improves CO2 physisorption capacity but also provides abundant reactive sites for photogenerated charge carriers. The increased density of surface active sites effectively captures and utilizes charge carriers, substantially reducing electron-hole recombination. Furthermore, the hierarchical pore architecture shortens diffusion pathways for reactants through localized concentration gradients, synergistically enhancing reaction efficiency via confined spatial effects and improved charge transport.
To elucidate the charge transfer properties of the photocatalysts, photoluminescence (PL) spectroscopy was employed to characterize CN-B and the CN-B/Ti3C2 (3:1) composite. The steady-state PL spectrum (Figure 6a) of CN-B exhibits a distinct emission peak at 450 nm, originating from the rapid recombination of photogenerated electrons and holes. In contrast, the CN-B/Ti3C2 (3:1) composite shows significantly quenched PL intensity, indicating that heterojunction formation establishes efficient charge transfer pathways, thereby suppressing non-radiative recombination. Time-resolved transient fluorescence decay spectra (Figure 6b) further reveal carrier dynamics. The average charge carrier lifetime of CN-B/Ti3C2 (3:1) is prolonged to 3.15 ns, notably longer than that of CN-B (2.07 ns). This prolonged lifetime confirms that the heterojunction facilitates directed electron transfer, enabling rapid migration of photogenerated electrons to surface reactive sites for CO2 reduction. The optimized charge separation mechanism provides a kinetic foundation for enhanced photocatalytic performance, allowing more high-energy electrons to participate in CO2 activation and conversion processes.
To further validate the enhanced charge separation in CN-B/Ti3C2 nanocomposites, transient photocurrent response and electrochemical impedance spectroscopy (EIS) measurements were conducted using a standard three-electrode system. As depicted in Figure 7a, the EIS Nyquist plots of CN-B/Ti3C2 (x:1) composites and CN-B—under both light and dark conditions—reveal significantly reduced arc radii for the composites compared to CN-B, indicating decreased charge transfer resistance. This reduction is attributed to facilitated charge separation and improved electrical conductivity resulting from effective heterojunction formation. Notably, the CN-B/Ti3C2 (4:1) sample exhibits increased resistance, which can be ascribed to excessive CN-B content leading to reduced conductivity, weakened heterojunction effects, impeded charge transport pathways, and increased interfacial defects. Among all composites, CN-B/Ti3C2 (3:1) demonstrates the smallest arc radius, confirming optimal charge transfer characteristics at the electrode/electrolyte interface. The high impedance observed under dark conditions in EIS measurements is characteristic of the semiconducting CN-B material, reflecting its limited charge-carrier generation in the absence of light. Photocurrent was measured without applied bias to evaluate the intrinsic photocatalytic activity under simulated solar irradiation, mimicking realistic solar-driven conditions. The photocurrent response under repeated on/off illumination cycles (Figure 7b) shows that the CN-B/Ti3C2 (3:1) electrode generates the highest photocurrent density, indicative of superior light absorption and charge separation efficiency. The reproducible photocurrent during cycling further underscores the catalyst’s excellent operational stability. In summary, the CN-B/Ti3C2 heterojunction catalyst exhibits enhanced light absorption, a narrower bandgap, improved charge separation, and superior charge transport properties compared to CN-B, collectively contributing to its outstanding photocatalytic CO2 reduction performance.

2.3. Photocatalytic Activity

The photocatalytic CO2 reduction activity of the samples was evaluated under simulated solar irradiation (Xe lamp) without any sacrificial agents, following the experimental details provided in Section 3.3. As illustrated in Figure 8a,b, pristine CN-B exhibited a CO production rate of 27.56 μmol·g−1·h−1. Multilayer Ti3C2 nanosheets alone showed limited activity, generating CO and CH4 at rates of 12.16 μmol·g−1·h−1 and 5.01 μmol·g−1·h−1, respectively. In the CN-B/Ti3C2 nanocomposite series, the CO2 reduction performance increased with the CN-B-to-Ti3C2 mass ratio up to 3:1. The optimal composite, CN-B/Ti3C2 (3:1), achieved a total CO yield of 320.84 μmol·g−1 over 4 h, along with a CH4 yield of 140.2 μmol·g−1. The CO production rate of this composite was 2.9 times higher than that of pure CN-B (110.24 μmol·g−1 under identical conditions). Further increasing the CN-B ratio to 4:1 resulted in decreased activity, due to reduced electrical conductivity [39], impaired charge separation, the aggregation of CN-B, and decreased active site availability.
Notably, the CN-B/Ti3C2 (3:1) composite maintained high and stable production rates of both CO and CH4 during a 48 h longevity test (Figure 8c). This consistent performance is attributed to the chemically robust heterojunction interface between CN-B and Ti3C2, which promotes structural integrity and sustained catalytic activity throughout prolonged photoreaction. Figure 8d displays the product selectivity distribution during the photocatalytic CO2 reduction over different catalysts. The CN-B/Ti3C2 (3:1) composite exhibits a slightly higher selectivity toward CH4 compared to other catalysts, indicating its enhanced ability to drive the multi-electron reduction pathway. As illustrated in Figure 8e, the apparent quantum yield (AQY) of CN-B/Ti3C2 (3:1) for CO2 reduction was measured at designated wavelengths (please refer to Table 1 for more detailed data), showing a gradual decrease in efficiency as the wavelength increases, which is consistent with the typical photo-response behavior of semiconductor-based photocatalysts. Furthermore, XRD analysis was performed on the CN-B/Ti3C2 (3:1) sample after a 48 h cycling test and compared with the pristine material (Figure 8f). The results reveal no significant changes in the crystal structure of the post-reaction sample, demonstrating the robust durability of the heterojunction system under prolonged photocatalytic conditions.
The CO2 reduction performance of the catalysts was evaluated under simulated solar irradiation (300 W Xe lamp) without sacrificial agents or photosensitizers. As shown in Figure 9a, in the presence of water vapor and without external temperature control, the CN-B/Ti3C2 (3:1) composite exhibited CO and CH4 production rates of 80.2 μmol·g−1·h−1 and 35.05 μmol g−1 h−1, respectively—2.9 times higher than that of CN-B (27.2 μmol g−1 h−1). However, when the temperature was maintained at 20 °C using a circulator chiller, the activity decreased significantly for all samples (Figure 9b), underscoring the important role of photothermal heating in reaction kinetics. Notably, while the thermal effect contributes to the reaction kinetics by providing localized heating, it alone is not the primary driver of the performance enhancement; instead, photocatalysis remains the dominant pathway, with the photothermal effect acting synergistically to reduce the activation energy barriers for CO2 dissociation and intermediate formation. This synergy arises from the integration of photoexcited charge carriers—which drive the redox reactions—with localized thermal energy that weakens C=O bonds and accelerates surface processes, thereby optimizing the overall reaction efficiency without supplanting the essential role of light-induced charge separation (Figure 9d,e).
Control experiments were conducted to confirm the origin of CO and CH4 (Figure 9c). No products were detected when CO2 was replaced by N2, when water vapor was omitted, in the absence of catalyst, or without light irradiation, confirming that both CO2 and H2O are essential reactants and that the process is photo-driven [40,41].
The photothermal effect was further investigated by monitoring the surface temperature under 300 mW·cm−2 irradiation (Figure 9f–h). The temperature of CN-B/Ti3C2 (3:1) rapidly increased from 22.7 °C to 233.9 °C within 150 s, substantially higher than that of CN-B (reaching 26.0 °C from 131.4 °C). This enhanced photothermal conversion stems from the composite’s black color and broad-spectrum absorption, which efficiently converts photon energy into heat via lattice vibrations, thereby accelerating reaction dynamics [15].
The heterojunction formed between CN-B and Ti3C2 facilitates charge separation through a built-in electric field: electrons migrate to Ti3C2, while holes remain in CN-B [42]. This not only suppresses recombination but also releases additional energy during charge transfer, further enhancing photothermal effects. The synergy between photocatalysis and photothermal conversion significantly improves overall CO2 reduction efficiency.

2.4. Mechanism of CO2 Photoreduction

Figure 10a presents the in situ FTIR spectra of CO2 adsorption on CN-B/Ti3C2. The gradually increasing peaks at 2359 cm−1, corresponding to C=O stretching vibrations, confirm the adsorption of CO2 on the catalyst surface [43]. Using the adsorption equilibrium state as the background, the dynamic interfacial interactions during photocatalysis are revealed in Figure 10b. Under 15 min of illumination, multiple characteristic absorption bands emerge: bicarbonate (HCO3; δ(C=O-H+) at 1067 cm−1), monodentate carbonate (m-CO32−; ν(C-O) at 1337 cm−1), and bidentate carbonate (b-CO32−; ν(C-O) at 1266 cm−1), and the peaks at 1430 cm−1 were considered to be the CO32−, all originating from CO2 hydration [44,45]. Additionally, A nascent positive peak emerged at 2368 cm−1, adjacent to the declining CO2 stretching vibrations, suggesting the formation of a protonated CO2 adspecies (O=C=O-H+), and absorption band at 1511 cm−1 is attributed to carboxylate (COOH*) species, indicating the formation of key intermediates via reaction between adsorbed CO2 and H+ [46]. The consumption of water and CO2 is evidenced by peaks at 1634 cm−1 and 2338 cm−1, highlighting the essential role of H2O in the reduction process [47]. These results collectively demonstrate that CN-B/Ti3C2 facilitates CO2 reduction through efficient adsorption, proton-assisted activation, and hydrolysis-mediated intermediate formation.
Based on the above results and analysis, Figure 11 illustrates the proposed mechanism for photocatalytic CO2 reduction over the CN-B/Ti3C2 (3:1) composite. Under light irradiation, the CN-B/Ti3C2 heterostructure functions as a Schottky junction, where metallic Ti3C2 forms a directional electron-extraction interface with semiconducting CN-B, markedly enhancing charge separation and photothermal-coupled catalytic activity [48]. The difference in Fermi levels between CN-B and Ti3C2 drives electron transfer from CN-B to Ti3C2, establishing a built-in electric field that promotes the spatial separation of photogenerated carriers: electrons migrate from the conduction band of CN-B to Ti3C2, while holes move from the valence band of Ti3C2 to CN-B. This directed charge flow significantly reduces recombination probability. The high electrical conductivity of Ti3C2 enables rapid electron transport to surface active sites, where CO2 is reduced stepwise to CO and CH4. Meanwhile, holes accumulated in the valence band of CN-B oxidize water molecules, facilitating proton generation and O2 evolution [49].
In situ infrared spectroscopy under illumination identifies key intermediate species, including adsorbed CO2* (bent at ~2359 cm−1) and COOH* (1710 cm−1), confirming efficient CO2 activation and proton-coupled reduction pathways. The overall process involves: (i) photoexcitation and thermal effects, where visible and near-infrared light generate electron–hole pairs while photothermal heating (enhanced by the black matrix and Ti3C2 plasmonics) elevates the local temperature, promoting hot carrier generation and weakening C=O bonds; (ii) efficient charge dynamics across the Schottky barrier, with electrons rapidly injected into Ti3C2 and suppressed recombination; and (iii) surface reduction processes wherein CO2 is converted via •CO2 → COOH* → CO* → CO/CH4, with protons supplied through water oxidation [46,50]. The intimate interfacial contact and synergistic electronic–thermal effects enable continuous solar-driven CO2 reduction, positioning CN-B/Ti3C2 as an advanced system for efficient and stable solar-to-chemical energy conversion. The CO2 photoreduction process over CN-B/Ti3C2 can be inferred as follows:
CN-B/Ti3C2 + hv → CN-B/Ti3C2 (e + h+)
CO2 → CO2*
CO2* + e + H+ → COOH*
COOH* + e + H+ → CO* + H2O
CO* → CO (g) & CO* + 6e + 6H+ → CH4 (g) + H2O
H2O + h+ → OH* + H+
OH* + h+ → 1/2 O2 (g) + H+

3. Experimental Section

3.1. Methods

3.1.1. Materials and Chemicals

All reagents were used as received without further purification: urea, phloxine B, absolute ethanol, potassium ferricyanide, potassium ferrocyanide, oleic acid, polyvinylpyrrolidone (PVP), and barium sulfate (BaSO4) of analytical reagent (AR) grade were supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Potassium bromide (KBr) and sodium sulfate (Na2SO4) (AR grade) were purchased from Aladdin Industrial Co. (Shanghai, China). Hydrofluoric acid (HF, 40%, AR grade) and titanium aluminum carbide (Ti3AlC2, 98%, 200 mesh) were obtained from Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China) with HF handled in PTFE containers under strict safety protocols due to its high toxicity.

3.1.2. Synthesis of CN-B

CN-B photocatalysts were synthesized via a one-step calcination method. Briefly, 10 g of urea and xmg of phloxine B (x = 25, 35, 45) were homogenized by mortar grinding for 30 min. The mixture was transferred to an alumina crucible and heated in static air at 5 °C min−1 to 540 °C, followed by a 2 h isothermal treatment. After cooling to room temperature naturally, the product was alternately washed with deionized water and absolute ethanol via centrifugation (8000 rpm, 10 min, three cycles) to remove impurities. The resulting solid was dried at 60 °C for 12 h, yielding CN-B powders. Samples were labeled CN-B-0.025, CN-B-0.035, and CN-B-0.045 based on phloxine B mass ratios (0.25–0.45 wt%). For comparison, pristine g-C3N4 (denoted CN) was synthesized identically without phloxine B addition [26].

3.1.3. Synthesis of Ti3C2 MXene

Multilayered Ti3C2 nanosheets were synthesized via selective etching of aluminum from Ti3AlC2 MAX phase. Briefly, 2 g of Ti3AlC2 powder (200 mesh, 98%) was slowly added to 40 mL of 40 wt% hydrofluoric acid (HF) in a PTFE container over 10 min to prevent localized overheating. The mixture was maintained at 35 °C under vigorous magnetic stirring (800 rpm) for 24 h to ensure complete Al removal. The resulting suspension was centrifuged at 3500 rpm for 10 min, followed by repeated washing cycles with deionized water until the supernatant reached pH ≈ 7. The collected sediment was vacuum-dried at 60 °C for 12 h, yielding accordion-like Ti3C2 MXene [27].

3.1.4. Synthesis of CN-B/Ti3C2 Composites

CN-B/Ti3C2 heterojunctions were fabricated via a hydrothermal self-assembly strategy. First, 0.1 g CN-B powder was dispersed in 100 mL deionized water under sonication (40 kHz, 20 min) to form a homogeneous suspension. Subsequently, 0.1 g Ti3C2 MXene was added, followed by additional sonication (30 min) and magnetic stirring (600 rpm, 6 h) to ensure interfacial contact. The mixture was transferred to a 150 mL Teflon-lined autoclave and hydrothermally treated at 120 °C for 12 h. After natural cooling, the product was collected by centrifugation (3500 rpm, 10 min), washed alternately with deionized water and ethanol (3–5 cycles), and vacuum-dried at 60 °C for 12 h. To optimize the heterojunction, composites with varying CN-B:Ti3C2 mass ratios (1:1, 2:1, 3:1, 4:1) were synthesized by adjusting CN-B mass (0.1–0.4 g) while maintaining constant Ti3C2 mass (0.1 g). The resulting materials were labeled CN-B/Ti3C2 (n:1), where n denotes the CN-B:Ti3C2 mass ratio.

3.2. Characterizations

The XRD patterns were performed on a Shimazu-6100 powder X-ray diffractometer (Shimadzu, Kyoto, Japan) with Cu Kα radiation. At 2θ, the scanning range was 10–80° and the scanning speed 10°/min. The XPS data was obtained using a Shimadzu/Krayos AXIS Ultra DLD (Shimadzu, Kyoto, Japan). The HRTEM images were taken on a field emission electron microscope (Tecnai G2 F20, FEI Company, Hillsboro, OR, USA) with an acceleration voltage. The scanning electron microscopy (SEM) images were characterized on a field emission scanning electron microscope (FE-SEM, JSM-7001F, JEOL, Tokyo, Japan). The ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS) was carried out on a Shimadzu UV-3600 spectrometer (Shimadzu, Kyoto, Japan). The photoluminescence (PL) spectra for solid power were investigated by an F4500 (Hitachi, Tokyo, Japan) and the xenon (Xe) lamp with an excitation wavelength of 375 nm. In situ Fourier transform infrared spectroscopy (in situ FTIR) was carried out to research the CO2 photoreduction process (Thermo Fisher Nicolet iS-10, Thermo Fisher Scientific, Waltham, MA USA). Thermal imaging temperature photos were obtained using an Infrared Thermal Imagers (FLIR Systems, Wilsonville, OR, USA). Photoelectrochemistry experiments including electrochemical impedance spectroscopy (EIS), and transient photocurrent responses (TPR) were studied.
The XRD patterns were performed on a Shimadzu-6100 powder X-ray diffractometer (Shimadzu, Kyoto, Japan) with Cu Kα radiation. At 2θ, the scanning range was 10–80° and the scanning speed 10°/min. The XPS data was obtained using a Shimadzu/Krayos AXIS Ultra DLD (Shimadzu, Kyoto, Japan). The HRTEM images were taken on a field emission electron microscope (Tecnai G2 F20, FEI Company, Hillsboro, OR, USA) with an acceleration voltage. The scanning electron microscopy (SEM) images were characterized on a field emission scanning electron microscope (FE-SEM, JSM-7001F, JEOL, Tokyo, Japan). The ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS) was carried out on a Shimadzu UV-3600 spectrometer (Shimadzu, Kyoto, Japan). The photoluminescence (PL) spectra for solid powder were investigated by an F4500 spectrometer (Hitachi, Tokyo, Japan) using a xenon (Xe) lamp with an excitation wavelength of 375 nm. In situ Fourier transform infrared spectroscopy (in situ FTIR) was carried out to research the CO2; photoreduction process using a Thermo Fisher Nicolet iS-10 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). Thermal imaging temperature photos were obtained using an Infrared Thermal Imager (FLIR Systems, Wilsonville, OR, USA). Photoelectrochemistry experiments including electrochemical impedance spectroscopy (EIS) and transient photocurrent responses (TPR) were studied.

3.3. CO2 Photoreduction Experiments

The CO2 photoreduction is conducted in a sealed 100 mL quartz homemade reactor using a 300 W Xe lamp (1000 mW cm−2, Beijing China Education Au-light Co., Ltd., Beijing, China) as the white light source. In a typical procedure, 10 mg of catalyst was dispersed in 3 mL of deionized water. The resulting mixture was dropped on quartz glass (area of 4 cm2) and then dried at 60 °C. The dried catalyst was placed in the quartz reactor with 1 mL of H2O. CO2 was then introduced into the reactor for 30 min to completely remove the air Gas products were detected by a gas chromatography (GC 5890N, Agilent Technologies, Santa Clara, CA, USA) equipped with hydrogen flame ionization detector (FID, Agilent Technologies, Santa Clara, CA, USA).

3.4. Photo-Electrochemical Measurements

The preparation of working electrodes began by mechanically grinding the CN-B and CN-B/Ti3C2 (x:1) powders into fine particles. Subsequently, 0.05 g of each powdered sample was uniformly dispersed in separate 3.0 mL ethanol solutions. To this dispersion, 0.03 mL of oleic acid and 0.01 g of polyvinylpyridone (PVP) were added as modifiers to enhance electrode film formation. The resulting homogeneous mixture was spin-coated onto a 1 × 1 cm2 fluorine-doped tin oxide (FTO) glass substrate, followed by air-drying at ambient temperature to form the working electrode. For electrochemical characterization, a conventional three-electrode setup was employed, consisting of a platinum (Pt) wire counter electrode, a saturated calomel electrode (SCE) as the reference, and 0.5 M Na2SO4 aqueous solution as the electrolyte. Electrochemical measurements were performed using a Chen Hua electrochemical workstation, applying a 5 mV sinusoidal potential waveform. All resulting data were recorded and analyzed using a CHI660 B electrochemical analyzer (Chen Hua Instruments, Shanghai, China).

3.5. Calculation of AQYs Parameters for Photoreduction of Carbon Dioxide by CN-B/Ti3C2 (3:1) at Different Wavelengths

When wavelength λ = 380 nm (CN-B/Ti3C2 (3:1)):
The number of incident photons:
N = E λ h c = 20 × 4 × 10 3 × 1 × 3600 × 450 × 10 9 6.626 × 10 34 × 3 × 10 8 = 6.52 × 10 20
AQY:
A Q Y = 8 × t h e   n u m b e r   o f   e v o l v e d   C H 4   m o l e c u l e s N = 8 × 6.02 × 10 23 × 0.447 × 10 6 6.52 × 10 20 = 0.33 %

4. Conclusions

In this study, we successfully synthesized a series of CN-B/Ti3C2 (x:1) composite photocatalysts via a hydrothermal method. The optimal sample, CN-B/Ti3C2 (3:1), exhibited exceptional photocatalytic CO2 reduction performance without the use of cocatalysts, achieving remarkable CO and CH4 production rates of 80.21 μmol·g−1·h−1 and 35.13 μmol·g−1·h−1, respectively—2.9 times higher than that of pristine CN-B. Furthermore, the composite demonstrated outstanding stability, maintaining high activity over a 24 h cyclic test. The significantly enhanced performance is attributed to the synergistic effects of the well-designed heterojunction: efficient charge separation and transfer across the CN-B/Ti3C2 interface substantially reduces electron–hole recombination, while the high conductivity of Ti3C2 facilitates rapid electron migration and CN-B stabilizes holes, collectively boosting photocatalytic efficiency. Moreover, the composite exhibits superior photothermal conversion capability, rapidly reaching 233.9 °C under light irradiation. This photothermal effect provides additional thermal energy that lowers the activation barrier, promotes carrier generation and separation, and accelerates reaction kinetics. This work highlights the importance of integrating heterojunction engineering with photothermal effects for designing highly efficient and stable photocatalysts. The CN-B/Ti3C2 composite represents a promising candidate for solar-driven CO2 conversion, offering valuable insights for developing advanced catalytic systems toward artificial carbon cycling and renewable energy utilization.

Author Contributions

Investigation, data curation, and writing of the original draft, M.C. and S.L.; resources, W.P.U.; writing—review, Y.Y. and J.Z.; conceptualization, supervision, and writing—review and editing, Z.Z. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 22208127). The Senior Talent Research Foundation of Jiangsu University (No. 23JDG030, 22JDG017), and the RGC Postdoctoral Fellowship Scheme of Hong Kong (RGC-PDFS-2324-2S04).

Data Availability Statement

All data are available as download within the database.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Abram, N.; Purich, A.; England, M.; McCormack, F.; Strugnell, J.; Bergstrom, D.; Vance, T.; Stål, T.; Wienecke, B.; Heil, P.; et al. Emerging evidence of abrupt changes in the Antarctic environment. Nature 2025, 644, 621–633. [Google Scholar] [CrossRef] [PubMed]
  2. Khan, K.; Tang, Y.; Cheng, P.; Song, Y.; Li, X.; Lou, J.; Iqbal, B.; Zhao, X.; Hameed, R.; Li, G.; et al. Effects of degradable and non-degradable microplastics and oxytetracycline co-exposure on soil N2O and CO2 emissions. Appl. Soil Ecol. 2024, 197, 105331. [Google Scholar] [CrossRef]
  3. Zhu, Z.; Xing, X.; Qi, Q.; Li, H.; Han, D.; Song, X.; Tang, X.; Ng, Y.; Huo, P. Regulation CN reduction of CO2 products selectivity by adjusting the number of V sites and mechanism exploration. Fuel 2025, 388, 134509. [Google Scholar] [CrossRef]
  4. Ozin, G. Accelerated optochemical engineering solutions to CO2 photocatalysis for a sustainable future. Matter 2022, 5, 2594. [Google Scholar] [CrossRef]
  5. Yang, J.; Ma, B.; Zhu, Y. A novel NDI/PTA organic heterojunction photocatalyst with built-in electric field for efficient hydrogen production. Appl. Catal. B Environ. Energy 2026, 381, 125890. [Google Scholar] [CrossRef]
  6. Bi, X.; Li, L.; Luo, L.; Liu, X.; Li, J.; You, T. A ratiometric fluorescence aptasensor based on photoinduced electron transfer from CdTe QDs to WS2 NTs for the sensitive detection of zearalenone in cereal crops. Food Chem. 2022, 385, 132657. [Google Scholar] [CrossRef] [PubMed]
  7. Hou, S.; Gao, X.; Lv, X.; Zhao, Y.; Yin, X.; Liu, Y.; Fang, J.; Yu, X.; Ma, X.; Ma, T.; et al. Decade Milestone Advancement of Defect-Engineered g-C3N4 for Solar Catalytic Applications. Nano-Micro Lett. 2024, 16, 70. [Google Scholar] [CrossRef]
  8. Li, B.; Li, N.; Guo, Z.; Wang, C.; Zhu, Z.; Tang, X.; Andrew Penyikie, G.; Wang, X.; Huo, P. Understanding the effect of multiple configurations in polymeric carbon nitride for efficient solar-driven hydrogen peroxide photosynthesis. Chem. Eng. J. 2024, 491, 152022. [Google Scholar] [CrossRef]
  9. Liu, Z.; Wang, L.; Liu, P.; Zhao, K.; Ye, S.; Liang, G. Rapid, ultrasensitive and non-enzyme electrochemiluminescence detection of hydrogen peroxide in food based on the ssDNA/g-C3N4 nanosheets hybrid. Food Chem. 2021, 357, 129753. [Google Scholar] [CrossRef]
  10. Wang, Z.; Yang, Z.; Fang, R.; Yan, Y.; Ran, J.; Zhang, L. A State-of-the-art review on action mechanism of photothermal catalytic reduction of CO2 in full solar spectrum. Chem. Eng. J. 2022, 429, 132322. [Google Scholar] [CrossRef]
  11. Lin, Y.; Yang, F.; Zhong, L.; Yu, D. Photoelectric/Photothermal Synergy Activation in Floating Aerogel Cathode Breaks Neutral Triphase Reaction Barriers for High-Efficiency Full-Spectrum-Responsive Seawater Batteries. Angew. Chem. Int. Ed. 2025, 64, e202508644. [Google Scholar] [CrossRef]
  12. He, H.; Ren, Y.; Zhu, Y.; Peng, R.; Lan, S.; Zhou, J.; Yang, B.; Si, Y.; Li, N. Continuous Flow Photothermal Catalytic CO2 Reduction: Materials, Mechanisms, and System Design. ACS Catal. 2025, 15, 10480–10520. [Google Scholar] [CrossRef]
  13. Kuang, L.; Chen, Z.; Yan, Y.; Guo, F.; Shi, W. Research progress of g-C3N4-based materials for photothermal-assisted photocatalysis. Int. J. Hydrogen Energy 2024, 87, 20–49. [Google Scholar] [CrossRef]
  14. Lei, D.; Su, D.; Maier, S. New insights into plasmonic hot-electron dynamics. Light Sci. Appl. 2024, 13, 243. [Google Scholar]
  15. Chen, Y.; Lin, X.; Li, W.; Sun, H.; Jia, S.; Zhou, Y.; Hao, Y.; Liu, Z.; Yin, S.; Guo, C.; et al. Harnessing Photo-to-Thermal Conversion in Sulfur-Vulcanized Mxene for High-Efficiency Solar-to-Carbon-Fuel Synthesis. Adv. Funct. Mater. 2024, 34, 2400121. [Google Scholar] [CrossRef]
  16. Karmakar, S.; Barman, S.; Rahimi, F.; Rambabu, D.; Nath, S.; Maji, T. Confining charge-transfer complex in a metal-organic framework for photocatalytic CO2 reduction in water. Nat. Commun. 2023, 14, 4508. [Google Scholar] [CrossRef]
  17. Han, M.; Yin, X.; Wu, H.; Hou, Z.; Song, C.; Li, X.; Zhang, L.; Cheng, L. Ti3C2 MXenes with Modified Surface for High-Performance Electromagnetic Absorption and Shielding in the X-Band. ACS Appl. Mater. Interfaces 2016, 8, 21011. [Google Scholar] [CrossRef]
  18. Du, X.; Du, W.; Sun, J.; Jiang, D. Self-powered photoelectrochemical sensor for chlorpyrifos detection in fruit and vegetables based on metal-ligand charge transfer effect by Ti3C2 based Schottky junction. Food Chem. 2022, 385, 132731. [Google Scholar] [CrossRef] [PubMed]
  19. Faraji, M.; Bafekry, A.; Fadlallah, M.; Molaei, F.; Hieu, N.; Qian, P.; Ghergherehchi, M.; Gogova, D. Surface modification of titanium carbide MXene monolayers (Ti2C and Ti3C2) via chalcogenide and halogenide atoms. Phys. Chem. Chem. Phys. 2021, 23, 15319. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, H.; Tang, Q.; Wu, Z. Construction of Few-Layer Ti3C2 MXene and Boron-Doped g-C3N4 for Enhanced Photocatalytic CO2 Reduction. ACS Sustain. Chem. Eng. 2021, 9, 8425–8434. [Google Scholar] [CrossRef]
  21. Liu, X.; Mu, Y.; Shi, L.; Yu, J.; Tao, R.; Chu, Z.; Liu, S.; Fan, X.; Liu, K. Enhanced photocatalytic overall water splitting via Al-doped SrTiO3/Ti3C2 Mxene Schottky heterojunction. Int. J. Hydrogen Energy 2025, 170, 151173. [Google Scholar] [CrossRef]
  22. Xiao, W.; Yu, H.; Xu, C.; Pu, Z.; Cheng, X.; Yu, F.; Liu, C.; Zhang, Q.; Zou, Z. 2D/2D Ti3C2 MXene/HTiNbO5 nanosheets Schottky heterojunction for boosting photothermal-assisted solar-driven photodegradation of tetracycline hydrochloride. J. Mater. Sci. Technol. 2024, 180, 193–206. [Google Scholar] [CrossRef]
  23. Xing, W.; Lin, H.; Shao, W.; Jian, N.; Liu, T.; Wen, S.; Han, J.; Wu, G. Heteroatom- and bonded tailored Schottky barrier in Ti3C2 MXene/carbon nitride heterojunctions toward efficient photocatalysis. Chem. Eng. J. 2025, 520, 166126. [Google Scholar] [CrossRef]
  24. Han, Y.; Zhang, Y.; Ye, L.; Gong, T.; Lu, X.; Yan, N.; Fu, Y. Establishing Dual-Interface Built-InElectric Fieldswithin Janus Heterostructures for Cooperative Photoredox Catalysis. J. Am. Chem. Soc. 2025, 147, 18637–18650. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Y.; Ali, A.; Osman, H.; Samad, S.; Abdulrahman, A.; Zhang, K.; Mahariq, I.; Kutliev, U.; Abduvokhidov, A. Engineering a Multifunctional Cu2O/BiVO4@Ti3C2 MXene Z-Scheme heterojunction for photocatalytic environmental Detoxification, H2 Generation, and CO2 reduction to CO and CH4: Experimental modeling and mechanistic insights. Sep. Purif. Technol. 2025, 369, 133110. [Google Scholar] [CrossRef]
  26. Kuate, L.; Chen, Z.; Lu, J.; Wen, H.; Guo, F.; Shi, W. Photothermal-Assisted Photocatalytic Degradation of Tetracycline in Seawater Based on the Black g-C3N4 Nanosheets with Cyano Group Defects. Catalysts 2023, 13, 1147. [Google Scholar] [CrossRef]
  27. Wang, X.; Li, H.; Li, H.; Lin, S.; Ding, W.; Zhu, X.; Sheng, Z.; Wang, H.; Zhu, X.; Sun, Y. 2D/2D 1T-MoS2/Ti3C2 MXene Heterostructure withExcellent Supercapacitor Performance. Adv. Funct. Mater. 2020, 30, 1910302. [Google Scholar]
  28. Saini, B.; Harikrishna, K.; Laishram, D.; Krishnapriya, R.; Singhal, R.; Sharma, R. Roleof ZnO in ZnO Nanoflake/Ti3C2 MXene Compositesin Photocatalyticand Electrocatalytic Hydrogen Evolution. ACS Appl. Nano Mater. 2022, 5, 9319–9333. [Google Scholar] [CrossRef]
  29. Zhang, L.; Yang, J.; Xie, T.; Feng, S.; Xue, L. Boosting visible-light-driven photocatalytic activity of BiPO4 via constructing Schottky junction with Ti3C2 MXene. Mater. Des. 2020, 192, 108772. [Google Scholar] [CrossRef]
  30. Shi, H.; Shao, R.; Lei, S.; Zhou, M.; Yuan, X.; Chen, Z.; Wei, S.; Zhong, S.; Zhao, Y.; Zhao, L.; et al. Breaking Janus shells in hollow heterostructured photocatalysts with spatially separated redox sites for enhanced synergy of CO2 methanation and H2O oxidation. Appl. Catal. B Environ. Energy 2025, 373, 125363. [Google Scholar] [CrossRef]
  31. Luo, L.; Huang, Y.; Cheng, K.; Alhassan, A.; Alqahtani, M.; Tang, L.; Wu, J. MXene-GaN van der Waals metal-semiconductor junctions for high performance multiple quantum well photodetectors. Light-Sci. Appl. 2021, 10, 177. [Google Scholar] [CrossRef]
  32. Pacheco-Peña, V.; Hallam, T.; Healy, N. MXene supported surface plasmons on telecommunications optical fibers. Light-Sci. Appl. 2022, 11, 22. [Google Scholar] [CrossRef]
  33. Jiang, L.; Zheng, Y.; Zhao, W.; Li, D.; Liu, M.; Pan, W.; Lan, M.; Huang, M.; Zhou, Y.; Jin, S. Ti3C2/AuNR heterojunction for near-infrared II plasma enhanced photothermal and photodynamic synergistic antibacterial therapy. Chem. Eng. J. 2025, 518, 164804. [Google Scholar] [CrossRef]
  34. Shen, W.; Qi, Q.; Hu, B.; Zhu, Z.; Huo, P.; Jiang, J.; Tang, X. Studying bimetals copper-indium for enhancing PCN photocatalytic CO2 reduction activity and selectivity mechanism. J. Ind. Eng. Chem. 2025, 145, 384–394. [Google Scholar] [CrossRef]
  35. Jing, X.; Mi, X.; Lu, W.; Lu, N.; Du, S.; Wang, G.; Zhang, Z. Edge effect-enhanced CO2 adsorption and photo-reduction over g-C3N4 nanosheet. Chin. J. Catal. 2024, 67, 112–123. [Google Scholar] [CrossRef]
  36. Otgonbayar, Z.; Oh, W. Synthesis of 2D/2D structural Ti3C2 MXene/g-C3N4 via the Schottky junction with metal oxides: Photocatalytic CO2 reduction with a cationic scavenger. Appl. Mater. Today 2023, 32, 101814. [Google Scholar] [CrossRef]
  37. Li, J.; Peng, H.; Luo, B.; Cao, J.; Ma, L.; Jing, D. The enhanced photocatalytic and photothermal effects of Ti3C2 Mxene quantum dot/macrosco picporous graphitic carbon nitride heterojunction for Hydrogen Production. J. Colloid Interface Sci. 2023, 641, 309–318. [Google Scholar] [CrossRef]
  38. He, F.; Zhu, B.; Cheng, B.; Yu, J.; Ho, W.; Macyk, W. 2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity. Appl. Catal. B Environ. 2020, 272, 119006. [Google Scholar] [CrossRef]
  39. Saeed, A.; Chen, W.; Shah, A.; Zhang, Y.; Mehmood, I.; Liu, Y. Enhancement of photocatalytic CO2 reduction for novel Cd0.2Zn0.8S@Ti3C2 (MXenes) nanocomposites. J. CO2 Util. 2021, 47, 101501. [Google Scholar] [CrossRef]
  40. Zhou, Y.; Cai, J.; Sun, Y.; Jia, S.; Liu, Z.; Tang, X.; Hu, B.; Zhang, Y.; Yan, Y.; Zhu, Z. Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight. Catalysts 2024, 14, 546. [Google Scholar] [CrossRef]
  41. Deng, X.; Ke, Y.; Ding, J.; Zhou, Y.; Huang, H.; Liang, Q.; Kang, Z. Construction of ZnO@CDs@Co3O4 sandwich heterostructure with multi-interfacial electron-transfer toward enhanced photocatalytic CO2 reduction. Chin. Chem. Lett. 2024, 35, 109064. [Google Scholar] [CrossRef]
  42. Yu, J.; Zhang, H.; Liu, Q.; Yu, J.; Zhu, J.; Li, Y.; Li, R.; Wang, J. 2D/2D heterojunction of Ti3C2/porous few-layer g-C3N4 nanosheets for high-efficiency extraction of uranium (VI). Sep. Purif. Technol. 2023, 312, 123442. [Google Scholar] [CrossRef]
  43. Yan, J.; Sun, Y.; Cai, J.; Cai, M.; Hu, B.; Yan, Y.; Zhang, Y.; Tang, X. Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic CO2 Reduction and Mechanism Insight. Catalysts 2024, 14, 599. [Google Scholar] [CrossRef]
  44. Chen, X.; Cai, M.; Huo, P.; Yan, Y.; Zhang, Y.; Li, P.; Zhu, Z. Sulfur-vulcanized CoFe2O4 with high-efficiency photo-to-thermal conversion for enhanced CO2 reduction and mechanistic insights into selectivity. Carbon Capture Sci. Technol. 2025, 14, 100377. [Google Scholar] [CrossRef]
  45. Qi, Q.; Shen, W.; Cai, M.; Cai, J.; Hu, B.; Han, D.; Tang, X.; Zhu, Z.; Huo, P. Construction of Cu-Modified g-C3N4 Nanosheets for Photoinduced CO2 Reduction to CO and Selectivity Mechanism Insight. ACS Appl. Nano Mater. 2024, 7, 24788–24797. [Google Scholar] [CrossRef]
  46. Yin, S.; Zhou, Y.; Liu, Z.; Wang, H.; Zhao, X.; Zhu, Z.; Yan, Y.; Huo, P. Elucidating protonation pathways in CO2 photoreduction using the kinetic isotope effect. Nat. Commun. 2024, 15, 437. [Google Scholar] [CrossRef]
  47. Ma, M.; Fang, Y.; Huang, Z.; Wu, S.; He, W.; Ge, S.; Zheng, Z.; Zhou, Y.; Fa, W.; Wang, X. Mechanistic Insights Into H2O Dissociation in Overall Photo-/Electro-Catalytic CO2 Reduction. Angew. Chem. Int. Ed. 2025, 64, e202425195. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, H.; Liu, J.; Xiao, X.; Meng, H.; Wu, J.; Guo, C.; Zheng, M.; Wang, X.; Guo, S.; Jiang, B. Engineering of SnO2/TiO2 heterojunction compact interface with efficient charge transfer pathway for photocatalytic hydrogen evolution. Chin. Chem. Lett. 2023, 34, 107125. [Google Scholar] [CrossRef]
  49. Xu, H.; Xiao, R.; Huang, J.; Jiang, Y.; Zhao, C.; Yang, X. In situ construction of protonated g-C3N4/Ti3C2 MXene Schottky heterojunctions for efficient photocatalytic hydrogen production. Chin. J. Catal. 2021, 42, 107–114. [Google Scholar] [CrossRef]
  50. Li, L.; Chen, F.; Zhao, B.; Yu, Y. Understanding of the structural evolution of catalysts and identification of active species during CO2 conversion. Chin. Chem. Lett. 2024, 35, 109240. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) CN-B, (b) Ti3C2 and (c) CN-B/Ti3C2 (3:1), TEM images of (d) CN-B, (e) Ti3C2 and (f) CN-B/Ti3C2 (3:1), SEM image (g) and Mapping scan of CN-B/Ti3C2 (3:1), (h) C, (i) N, (j) Ti.
Figure 1. SEM images of (a) CN-B, (b) Ti3C2 and (c) CN-B/Ti3C2 (3:1), TEM images of (d) CN-B, (e) Ti3C2 and (f) CN-B/Ti3C2 (3:1), SEM image (g) and Mapping scan of CN-B/Ti3C2 (3:1), (h) C, (i) N, (j) Ti.
Catalysts 15 01037 g001
Figure 2. (a) XRD patterns of CN-B, Ti3C2 and CN-B/Ti3C2 (x:1) and (b) FT-IR spectrum.
Figure 2. (a) XRD patterns of CN-B, Ti3C2 and CN-B/Ti3C2 (x:1) and (b) FT-IR spectrum.
Catalysts 15 01037 g002
Figure 3. XPS spectra of CN-B and CN-B/Ti3C2 (3:1) catalyst. (a) Survey spectrum, Region XPS spectra of (b) C 1s, (c) N 1s and (d) Ti 2p.
Figure 3. XPS spectra of CN-B and CN-B/Ti3C2 (3:1) catalyst. (a) Survey spectrum, Region XPS spectra of (b) C 1s, (c) N 1s and (d) Ti 2p.
Catalysts 15 01037 g003
Figure 4. (a) UV-vis DRS of CN-B, Ti3C2 and CN-B/Ti3C2 (x:1); (b) Measured band gap values of CN-B and CN-B/Ti3C2 (3:1) samples; (c) XPS valence band spectra of the CN-B and CN-B/Ti3C2 (3:1); (d) electronic band structures of the CN-B and CN-B/Ti3C2 (3:1).
Figure 4. (a) UV-vis DRS of CN-B, Ti3C2 and CN-B/Ti3C2 (x:1); (b) Measured band gap values of CN-B and CN-B/Ti3C2 (3:1) samples; (c) XPS valence band spectra of the CN-B and CN-B/Ti3C2 (3:1); (d) electronic band structures of the CN-B and CN-B/Ti3C2 (3:1).
Catalysts 15 01037 g004
Figure 5. N2 adsorption–desorption curve (a) and pore distributions (b) of CN-B, Ti3C2 and CN-B/Ti3C2 (3:1).
Figure 5. N2 adsorption–desorption curve (a) and pore distributions (b) of CN-B, Ti3C2 and CN-B/Ti3C2 (3:1).
Catalysts 15 01037 g005
Figure 6. PL spectrum (a) and TS-PL curves (b) of CN-B and CN-B/Ti3C2 (3:1).
Figure 6. PL spectrum (a) and TS-PL curves (b) of CN-B and CN-B/Ti3C2 (3:1).
Catalysts 15 01037 g006
Figure 7. (a) EIS of CN-B and CN-B/Ti3C2 (x:1) sample and (b) Photocurrent curves.
Figure 7. (a) EIS of CN-B and CN-B/Ti3C2 (x:1) sample and (b) Photocurrent curves.
Catalysts 15 01037 g007
Figure 8. (a,b) Reaction rates for photocatalytic CO2 reduction in different samples. (c) twelve cycle experiments of photocatalytic carbon dioxide reduction on CN-B/Ti3C2 (3:1) photocatalyst. (d) Selectivity of the product of CN-B/Ti3C2 (3:1) in the reduction of CO2 reaction. (e) AQYs of photoreduction of CO2 by CN-B/Ti3C2 (3:1) at different wavelengths after 1 h irradiation. (f) Comparison of XRD of CN-B/Ti3C2 (3:1) after 48 h of carbon dioxide reduction and XRD of the sample without performance testing.
Figure 8. (a,b) Reaction rates for photocatalytic CO2 reduction in different samples. (c) twelve cycle experiments of photocatalytic carbon dioxide reduction on CN-B/Ti3C2 (3:1) photocatalyst. (d) Selectivity of the product of CN-B/Ti3C2 (3:1) in the reduction of CO2 reaction. (e) AQYs of photoreduction of CO2 by CN-B/Ti3C2 (3:1) at different wavelengths after 1 h irradiation. (f) Comparison of XRD of CN-B/Ti3C2 (3:1) after 48 h of carbon dioxide reduction and XRD of the sample without performance testing.
Catalysts 15 01037 g008
Figure 9. (a) CO yields of different catalysts without temperature control, (b) the same as (a) except that T = 293 K, (c) Performance of CN-B/Ti3C2 (3:1) under different reaction conditions, In the case of N2, no H2O, without catalyst, normal and no light, respectively. (d,e) Comparison of CO yield of CN-B/Ti3C2 (3:1) under different temperatures with and without white light, (f) The time curve of light and heat conversion of CN-B and CN-B/Ti3C2 (3:1) under xenon lamp (300 W, 1000 mW cm−2), (g) CN-B and (h) CN-B/Ti3C2 (3:1) infrared thermogram of the catalytic reaction in water vapor.
Figure 9. (a) CO yields of different catalysts without temperature control, (b) the same as (a) except that T = 293 K, (c) Performance of CN-B/Ti3C2 (3:1) under different reaction conditions, In the case of N2, no H2O, without catalyst, normal and no light, respectively. (d,e) Comparison of CO yield of CN-B/Ti3C2 (3:1) under different temperatures with and without white light, (f) The time curve of light and heat conversion of CN-B and CN-B/Ti3C2 (3:1) under xenon lamp (300 W, 1000 mW cm−2), (g) CN-B and (h) CN-B/Ti3C2 (3:1) infrared thermogram of the catalytic reaction in water vapor.
Catalysts 15 01037 g009
Figure 10. In situ FTIR spectra of CO2 adsorption (a) and reaction (b) of CN-B/Ti3C2 collected at different time intervals.
Figure 10. In situ FTIR spectra of CO2 adsorption (a) and reaction (b) of CN-B/Ti3C2 collected at different time intervals.
Catalysts 15 01037 g010
Figure 11. Schematic diagram of the mechanism of CN-B/Ti3C2 photocatalytic reduction of CO2.
Figure 11. Schematic diagram of the mechanism of CN-B/Ti3C2 photocatalytic reduction of CO2.
Catalysts 15 01037 g011
Table 1. Parameters to calculate AQYs of photoreduction CO2 over the CN-B/Ti3C2 (3:1) under different wavelengths.
Table 1. Parameters to calculate AQYs of photoreduction CO2 over the CN-B/Ti3C2 (3:1) under different wavelengths.
Wavelength (nm)CH4 Evolution
(μmol g−1)
Light Intensity
(mW cm−2)
Irradiation Time
(h)
Irradiation Area
(cm2)
AQY (%)
3800.44720140.33
4050.450520140.335
4600.248520140.185
5200.14220140.105
6000.105520140.08
7000.08820140.065
8500.062520140.045
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

Cai, M.; Lv, S.; Li, Y.; Utomo, W.P.; Yan, Y.; Zhu, Z.; Zhao, J. Interfacial Engineering of CN-B/Ti3C2 MXene Heterojunction for Synergistic Solar-Driven CO2 Reduction. Catalysts 2025, 15, 1037. https://doi.org/10.3390/catal15111037

AMA Style

Cai M, Lv S, Li Y, Utomo WP, Yan Y, Zhu Z, Zhao J. Interfacial Engineering of CN-B/Ti3C2 MXene Heterojunction for Synergistic Solar-Driven CO2 Reduction. Catalysts. 2025; 15(11):1037. https://doi.org/10.3390/catal15111037

Chicago/Turabian Style

Cai, Ming, Shaokun Lv, Yuanyuan Li, Wahyu Prasetyo Utomo, Yongsheng Yan, Zhi Zhu, and Jun Zhao. 2025. "Interfacial Engineering of CN-B/Ti3C2 MXene Heterojunction for Synergistic Solar-Driven CO2 Reduction" Catalysts 15, no. 11: 1037. https://doi.org/10.3390/catal15111037

APA Style

Cai, M., Lv, S., Li, Y., Utomo, W. P., Yan, Y., Zhu, Z., & Zhao, J. (2025). Interfacial Engineering of CN-B/Ti3C2 MXene Heterojunction for Synergistic Solar-Driven CO2 Reduction. Catalysts, 15(11), 1037. https://doi.org/10.3390/catal15111037

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