Next Article in Journal
Optimizing Metal Sites in Hierarchical USY for Selective Hydrocracking of Naphthalene to BTX
Previous Article in Journal
Natural Tyrosinase Inhibitors from Lycopodium japonicum
Previous Article in Special Issue
Understanding the Paradigm of Molecular-Network Conformations in Nanostructured Se-Rich Arsenoselenides AsxSe100−x (x < 10)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interface Catalysts of In Situ-Grown TiO2/MXenes for High-Faraday-Efficiency CO2 Reduction

1
U.S. Army Combat Capabilities Development Command Chemical Biological Center, Research & Operations Directorate, Aberdeen Proving Ground, Aberdeen, MD 21010, USA
2
Department of Chemistry and Biochemistry, Utah State University, Logan, UT 84322, USA
3
School of Aerospace and Mechanical Engineering, University of Oklahoma, Norman, OK 73019, USA
4
Department of Electrical and Computer Engineering, University of Delaware, Newark, DE 19711, USA
5
Department of Materials Science and Engineering, University of Delaware, Newark, DE 19711, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2025, 30(19), 4025; https://doi.org/10.3390/molecules30194025
Submission received: 12 August 2025 / Revised: 3 October 2025 / Accepted: 4 October 2025 / Published: 9 October 2025
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry, 3rd Edition)

Abstract

Climate change and the global energy crisis have led to an increasing need for greenhouse gas remediation and clean energy sources. The electrochemical CO2 reduction reaction (CO2RR) is a promising solution for both issues as it harvests waste CO2 and chemically reduces it to more useful forms. However, the high overpotential required for the reaction makes it electrochemically unfavorable. Here, we fabricate a novel electrode composed of TiO2 nanoparticles grown in situ on MXene charge acceptor 2D sheets with excellent CO2RR characteristics. A straightforward solvothermal method was used to grow the nanoparticles on the Ti3C2Tx MXene flakes. The electrochemical performance of the TiO2/MXene electrodes was analyzed. The Faradaic efficiencies of the TiO2/MXene electrodes were determined, with a value of 99.41% at −1.9 V (vs. Ag/AgCl). Density functional theory mechanistic analysis was used to reveal the most likely mechanism resulting in the production of one CO molecule along with a carbonate anion through ∗CO, ∗O, and activated CO22− intermediates. Bader charge analysis corroborated this pathway, showing that CO2 gains a greater negative charge when TiO2/MXene serves as a catalyst compared to MXene or TiO2 alone. These results show that TiO2/MXene nanocomposite electrodes may be very useful in the conversion of CO2 while still being efficient in both time and cost.

1. Introduction

Climate change and the global energy crisis have led researchers to turn to CO2 reduction as a means of greenhouse gas remediation to produce a clean energy source. The process of harvesting waste CO2 and chemically reducing it to a more useful form promises to address both issues but is energy-intensive. For this reason, efficient photo-, electro- and photoelectro-catalytic processes to promote the CO2 reduction reaction (CO2RR) are of great interest. The use of TiO2 as a catalyst is broad and multifaceted, which has led to many reviews on the process [1,2,3,4,5,6,7,8,9,10,11]. Of particular interest are the catalytic processes occurring at the TiO2 surfaces [5,10,11]. Over 50 years ago, TiO2 was demonstrated as a catalyst in water splitting [12]. By nature, the surface characteristics of a heterogeneous catalyst play a crucial role in its catalytic abilities. To this end, TiO2 surface structures and properties have been studied extensively with high-resolution scanning tunneling microscopy (STM) [13,14], scanning electron microscopy (SEM) [15,16], and transmission electron microscopy (TEM) [17,18]. TiO2 has found a role as a catalyst in the reduction of CO2 as well [19]. However, the catalytic reduction of CO2 to form methane or methanol on TiO2 nanoparticles is not without flaws: the close proximity of the reduction and oxidation half-reactions can lead to unwanted carrier recombination, and the reaction pathways proposed for the catalysis are complex and still debated [20,21,22]. As a potential solution, cocatalysts have been introduced to assist in the catalytic CO2RR by TiO2. Systems such as TiO2/Pt, MgO/Pt/TiO2, Ag/TiO2, and many more have been investigated for their catalytic potential in CO2RR and their variable performances have been cataloged [23]. One such cocatalyst, Ti3C2 (MXene), shows particular promise due to the high performance of the combined TiO2/Ti3C2 material and the absence of precious materials [24].
MXenes encompass a broad family of 2D layered materials composed of transition metal carbides and nitrides, with applications spanning from environmental remediation to photonics [25,26,27]. Specifically, Ti3C2Tx (T=F, O, or OH) was first fabricated in 2011 by Naguib et al. by selectively etching away the Al layer in MAX phase materials (Ti3AlC2) [28,29]. Although carbide MXenes take the general form of Tin+1CnTx, further references to MXene in this work refer specifically to Ti3C2Tx. MXenes have found uses in flexible light-emitting diodes (LEDs) [30], hydrogel sensors [31], brain activity recording [32], and adjustable-focus ocular lens devices [33]. Of particular interest are the energy storage capabilities of MXenes and their electrochemical uses [34,35,36,37]. Such charge storage capabilities have been ascribed to the layered structure of MXenes, where it has been shown that Na+ ions are reversibly transported into and out of the interlayer space resulting in expansion and contraction, respectively, in aqueous electrolytes [38]. Such intercalation of cations by MXenes has also been demonstrated for other species, both spontaneously and electrochemically, where the 2D sheets become capacitors [39]. In fact, MXene-based solid-state supercapacitors have been successfully deposited onto conductive polymer threads, advancing the usefulness of the material to a flexible system [40]. Due to these properties, MXene is an excellent host material for TiO2 photocatalysts acting as a charge acceptor and scaffold.
TiO2/MXene composite materials were developed some time ago and boasted increased performance over the individual components [41,42,43,44,45]. The first demonstration of a TiO2/MXene catalyst was for the photocatalytic CO2RR, and was achieved by calcining Ti3C2 [24]. This work monitored the CO2RR to methane and concluded that the mechanism involved the MXene scaffold preventing fast charge recombination at the TiO2 surfaces and emphasized that the in situ growth of TiO2 on MXene was crucial in enhancing charge separation. In the same year, a solvothermal process for the fabrication of TiO2/MXene was published and analyzed for its photocatalytic potential [46], finding that TiO2/MXene readily degraded carbamazepine through ·OH and ·O2 species. Electrocatalytically, TiO2/MXene composites were tested for their ability to reduce N2 to NH3, showing superior efficiency under acidic conditions and excellent durability [47]. In an interesting approach, TiO2 was married to a hybrid Ti3C2/Ru cocatalyst which provided enhanced charge separation that promoted the hydrogen evolution reaction (HER) [48]. Increased catalytic activity of TiO2/MXene over TiO2 in the oxidative dehydrogenation of ethane has also been observed, where the defects in the composite surface relative to TiO2 were shown to contribute to a 4× increase in performance [49]. With specific regard to CO2 reduction, attempts at designing TiO2/MXene-based materials have been made including C3N4/Ti3C2Tx/TiO2 [50,51], Ru-Ti3CN/TiO2 [52], TiO2/Ti3CN [53], and even tailoring the anatase/rutile moieties of TiO2 in TiO2A/Ti3C2/TiO2R MXene [54]. Due to the photocatalytic activity of TiO2, most of the work to date covering the catalytic CO2RR by TiO2/MXene has focused on the photocatalytic process. However, this has led to the direct electrocatalytic counterpart to be overlooked.
Here, we fabricate a novel electrode composed of TiO2 nanoparticles grown in situ on MXene charge acceptor 2D sheets with excellent electrocatalytic CO2RR characteristics. A straightforward solvothermal method was used to grow the nanoparticles on the Ti3C2Tx MXene flakes for electrochemical CO2RR. After thorough characterization of the composite material, we used experimental and theoretical analyses to hypothesize a mechanism to explain the high Faradaic efficiency of the CO2RR.

2. Results and Discussion

TiO2/MXene catalysts were fabricated as described below and illustrated in Figure 1. To characterize our samples, the XRD scattering patterns of TiO2/MXene, MXene, and pristine MAX are presented in Figure 2A. After etching, a shift of the (002) peak from 9.64° to 7.28° and the elimination of most intense peaks at 39.06° indicated removal of Al layers in MAX and the formation of exfoliated MXene [55]. After hydrothermal processing, there were two new peaks at 25.76° and 39.14° in TiO2/MXene, which correspond to the (101) and (004) facets of anatase TiO2 [56]. Furthermore, the peak shift from 7.28° to 6.42° indicated an increase in the interlayer distance from the exfoliation of TiO2/MXene after the solvothermal treatment.
Next, we investigated the chemical state and composition of TiO2/MXene and pristine MXene by X-ray photoelectron spectroscopy (XPS). The presence of Ti, O, C, F, and Cl is shown in Figures 2B–D, S2 and S3. The F and Cl elements present on the surface of MXene are a product of HF etching. For the Ti 2p region, the peaks at 458.3, 459.7, and 464.2 eV correspond to the Ti–O 2p3/2, Ti–F/Cl, and Ti–O 2p1/2, respectively. The C 1s region was fitted with three peaks at 284.6, 285.9, and 288.7 eV, ascribed to the C-C, C-O, and O-C-O bonds, respectively [6]. The O 1s region spectra can also be divided into three subpeaks at 529.9, 531.2, and 532.5 eV, which are ascribed to oxygen in TiO2 (Ti–O–Ti), oxygen in the surface hydroxyl group (Ti–O–H), and C–O–Ti, respectively. Compared with MXene [57], all signals in TiO2/MXene displayed negative shifts, indicating an increased electron transfer from MXene to TiO2. As a result, a strong interfacial interaction can be achieved at the TiO2/MXene interface, which may enhance the performance toward electrocatalyzing the CO2 reduction.
The morphology and microstructure of the synthesized TiO2/MXene were characterized by TEM and HRTEM. As shown in Figure 3A, the exfoliated MXene nanosheets showed single (or few)-layered and flake-like morphology using TEM. TiO2 nanoparticles with a size of 10–20 nm were also observed to be uniformly distributed on the surface of MXene nanosheets. As shown in Figure 3B, lattice fringes with a spacing of 0.244 nm were observed, which were ascribed to the (103) plane of anatase TiO2. Energy-dispersive X-ray (EDX) characterization was also conducted and the atomic ratio of Ti:O was found to be 2.96:1 (Figure S4). The elemental mapping results of the TiO2/MXene further confirm that the distribution of Ti and O elements of the in situ formation of TiO2 was monodisperse on the MXene surface, as shown in Figure 3C. TiO2 nanoparticles were grown in situ on MXene surfaces, endowing robustness to the microstructure and morphology.
The electrochemical conversion of CO2 to CO was investigated by cyclic voltammetry (CV) and controlled potential coulometry (CPC). Figure 4A shows the CVs collected using the TiO2/MXene electrocatalyst in the electrolyte saturated with N2 (black) or CO2 (red) gas. Here, we see a clear difference between the CVs upon introduction of CO2 during both increasing and decreasing potentials. In the N2-purged solution, TiO2/MXene showed a typical capacitive profile due to the CP support. The peak at −1.3 V is attributed to reduction from Ti4+ to Ti3+ on the TiO2 surface. With the CO2-saturated electrolyte, an apparent cathodic current was observed, corresponding to efficient CO2 conversion. Figure 4B shows the CPC scans of the N2-(black) or CO2-purged (red) electrolyte at a constant potential of −1.9 V. As such, the TiO2/MXene nanocomposites show significant potential-dependent CO2RR catalytic properties.
To investigate the performance of the TiO2/MXene catalyst during CO2RR, the Faradaic efficiency for CO production (FECO) was calculated for various potentials. As shown in Figure 4C, it was observed that CO production started at potentials more negative than −1.4 V and peaked at 99.41 ± 0.50% at −1.9 V. A further increase in potential led to a decrease in efficiency due to hydrogen evolution occurring at high potentials. It also is important to note that CO was not produced in the N2-saturated electrolyte (Figure S2), indicating that the generated CO originated from the electrocatalytic reduction of CO2, rather than from ACN or MXene decomposition.
To ensure that TiO2/MXene was responsible for CO2RR and not its components, we compared the FECO of TiO2/MXene, MXene, and commercial TiO2 at −1.9 V. As shown in Figure 4D, the TiO2/MXene electrode exhibits a near 100% FECO; however, almost no CO2 electroreduction resulted from electrodes fabricated using pure MXene or commercial TiO2. These results indicate the efficient CO2 reduction arises from the TiO2/MXene composite material. To confirm this hypothesis, the catalytic performance of the TiO2/MXene nanocomposite using different in situ growth times was studied (Figure S6). We found that increasing growth times up to 12 h produced higher FECO values, but excessive growth time reduced FECO. We hypothesize that the observed decrease in FECO was due to poor conductivity caused by the excess synthesis of TiO2 on MXene, as confirmed by the electrochemical impedance spectroscopy (EIS) results in Figure S7. To demonstrate the stability of our TiO2/MXene system, CO2RR was performed at increasing timescales from 30 min to 3 h. As depicted in Figure S8, after 3 h of electrocatalysis, FECO only decreased by 13%, demonstrating certain stability for potential high-throughput applications. Overall, our results suggest that the in situ growth of TiO2 on MXene sheets produces efficient catalytic sites for CO2RR with high Faradaic efficiency and selectivity, which is an improvement compared to the other non-precious electrocatalyst materials reported, such as Bi nanoparticles [58], TiS3 [59], and TiO2 [60], as shown in Table S1.
To best understand the CO2RR mechanism on TiO2/MXene, density functional theory (DFT) calculations were performed using the Vienna ab initio simulation package with the Perdew–Burke–Ernzerh functional [61,62,63,64]. The projector-augmented wave pseudopotentials were implemented for the exchange correlation function and pseudopotential treatment [65,66]. The structure of the TiO2/MXene used in these considerations is shown in Figure 5A. The calculations were conducted on three systems: (1) TiO2 nanoparticles, (2) MXene, and (3) TiO2/MXene. The kinetic cutoff energy was 600 eV throughout the simulation. The Brillouin zone was sampled by a Monkhorst–Pack grid centered at the γ point with a k-point mesh of 3 × 3 × 3 for TiO2 and 3 × 3 × 1 for MXene and TiO2/MXene. All structures analyzed by DFT were fully optimized, and computations were continued until they met the convergence thresholds of having a total energy of 10−6 eV and atomic forces not exceeding 0.05 eV/Å. Other model details, e.g., the sizes of MXene (112 atoms) and TiO2 nanoparticles (42 atoms), the construction of slab models, and the nanoparticles (Figure S9), are included in the Supplementary Materials. The size of systems used for modeling was selected as a balance between computational efficiency and accuracy [61]. The analysis suggests that the (100) direction of TiO2 nanoparticles is most likely to be perpendicular to the (110) direction of MXene (Figure S10). Furthermore, the strong affinity between the TiO2 nanoparticles and MXene is primarily governed by the interaction involving oxygen atoms of TiO2 and titanium atoms of MXene, facilitated through oxygen dangling bonds. The resulting TiO2/MXene model is illustrated in Figure 5A and serves as the system upon which subsequent calculations are based.
To perform the adsorption energy analysis of CO2 onto the TiO2/MXene, we examined five geometrically unique adsorption configurations, as shown in Figure 5B, where the negative shift in adsorption energy indicates that the process is favorable. Interestingly, the adsorption free energy magnitude is comparatively weaker than that of other catalytic systems, placing TiO2/MXene within the optimal range for catalytic adsorption of CO2 [67,68]. From these results, we can ascertain that the CO2 molecules are adsorbed above the 17th and 12th Ti atoms, which are selected for further analysis. It is worth noting that the pattern of adsorption energy does not exist when CO2 is adsorbed onto MXene or TiO2 (Figures S11 and S12), where the elevated adsorption energy could potentially poison the adsorbent.
A Gibbs free energy change analysis was conducted to understand the energetics of intermediates while the reaction proceeds. The analysis was performed following three possible reaction paths for CO2 reduction in an aprotic environment. Step (R1) is the adsorption of the first CO2 molecule and reaction. Step (R2) is the successive adsorption of the second CO2 molecule onto the catalytic system. The reactions given in steps (R3a–c) represent the possible reductive paths of the two adsorbed CO2 molecules. Figure 5C depicts the Gibbs free energy changes for reactants and intermediates present in the reaction on TiO2/MXene, compared with pure MXene and TiO2 nanoparticles.
+ C O 2 C O 2
C O 2 + C O 2
2 C O 2 C O + O + C O 2 C O + C O 3 2
2 C O 2 + 2 e 2 C O 2 2 o x a l a t e ( C 2 O 4 )
2 C O 2 2 C O + O 2
Our free energy analysis leads to three unique systems for the potential catalysts, as shown in Figure 5C. (1) In the case of pure MXene (brown lines), we observe a gradual change in free energy as CO2 molecules are successively adsorbed. This incremental change suggests that reductive reactions on pure MXene surfaces are less likely to occur. (2) In the case of pure TiO2 nanoparticles (orange lines), the free energy change undergoes a significant negative shift upon adsorption of the first and second CO2 molecules. These adsorption processes lead to the formation of an oxalate group, accompanied by a substantial negative change in free energy (−9.20 eV). This oxalate group eventually dissociates into two CO2 molecules. (3) The case of the TiO2/MXene catalytic system is more nuanced. Initially, there is a slight positive change in free energy (0.07 eV) upon the adsorption of the first CO2 molecule, which reduces it to 0.02 eV when the second CO2 molecule is adsorbed near the first. In this scenario, three possibilities arise: (R3a) the two dimerized CO2 molecules can form one oxalate, (R3b) two CO molecules and one O2 molecule may be generated, and R3c) the formation of one CO molecule along with a carbonate anion (CO32−) via intermediates such as ∗CO, ∗O, and activated CO22−. Of these options, the first and second cases exhibit positive changes in free energy (0.06 eV and 6.64 eV, respectively), making these reactions less favorable. However, in the third scenario, the formation of one CO molecule and one CO32− shows a negative change in free energy (−0.51 eV), indicating that the CO2RR is energetically favorable in this direction. Further support for these findings comes from the Bader charge analysis shown in Figure 5D, where the net charge on adsorbed CO2 molecules reveals that they gain a higher negative charge when TiO2/MXene serves as a catalyst compared to pure MXene or TiO2. The results of the Bader analysis are visualized in Figure 5E, where the yellow shading indicates areas of charge accumulation, and cyan shows those of charge depletion. This suggests a significant charge transfer from TiO2/MXene to CO2 molecules. In the future, other product analyses may be performed including in situ spectroscopy [69], and possibly interfacial intermediate analyses through sum-frequency generation spectroscopy. Additionally, while our TiO2/MXene electrocatalyst showed high FE for the CO2RR to CO, it is at the cost of a high overpotential of 1.9 V vs. Ag/AgCl. Future strategies to reduce this overpotential include composite materials such as TiI4 to modify the electronic structure and promote more marketable reaction conditions [70].

3. Experiment

Materials. Ti3AlC2 (MAX) powders were purchased from Forsman (Beijing, China). Carbon paper (HCP020N, CP) was purchased from Hesen (Shanghai, China). Nickel foam (NF) was provided by Hefei Kejing Materials Technology (Hefei, Anhui, China). Nafion@117 solution was purchased from Sigma-Aldrich (St. Louis, MO, USA). Acetonitrile was purchased from Fisher Chemical (Pittsburgh, PA, USA). Ethanol and hydrochloric acid were purchased from Pharmco (Brookfield, CT, USA). Lithium fluoride was purchased from Alfa Aesar (Ward Hill, MA, USA). Potassium hexafluorophosphate was purchased from Oakwood Chemical (Estill, SC, USA). Deionized (DI) (18.2 MΩ cm) water was produced by using a Milli-Q ultrapure system (Thermo Scientific Barnstead E-Pure, Waltham, MA, USA).
Preparation of MXene. Ti3C2Tx MXene was obtained via an HF etching method [71]. A total of 100 mg of LiF was added into 2 mL of 9 M HCl solution in an ice bath before 100 mg of MAX was added into the solution. After etching at 35 °C for 24 h, the obtained MXene was washed 8 times with DI water and centrifuged at 3000 rpm for 1 h. The collected MXene was freeze-dried and stored in a vacuum desiccator.
Preparation of TiO2/MXene. As illustrated in Figure 1, 20 mg of Ti3C2Tx MXene was dispersed in 2 mL ethanol and sonicated at 400 W for 2 h. The suspension was transferred into a 40 mL PTFE autoclave for solvothermal treatment at 120 °C for 12 h. Afterward, the as-prepared suspension was centrifuged at 3000 rpm for 20 min. The obtained solid was freeze-dried and stored in a vacuum desiccator.
Characterizations. X-ray diffraction (XRD) patterns were analyzed by using a Rigaku Miniflex II X-ray diffractometer (Rigaku Americas, The Woodlands, TX, USA) with Cu Kα radiation (λ = 1.5406 Å). Transmission electron microscopy (TEM), high-resolution transmission electron microscopy (HR-TEM) images, and elemental mapping analysis were collected using a FEI Tecnai G2 F20 FE-TEM (FEI, Hillsboro, OR, USA) microscopy system. X-ray photoelectron spectroscopy (XPS) measurements were performed using a Kratos Axis Ultra instrument (Kratos Analytical Ltd., Manchester, UK).
Preparation of the Working Electrode. A total of 4 mg of th TiO2/MXene catalyst was suspended in a mixture of 960 μL Di-water and 40 μL Nafion@117 solution, followed by a 2 h sonication to form a homogeneous ink. A total of 100 μL of the ink (0.4 mg of catalyst) was deposited onto a carbon paper strip with an area of 1 cm2 and dried in a vacuum desiccator.
Electrocatalytic CO2 Reduction. All potentials are reported with respect to the Ag/AgCl reference electrode. Electrochemical CO2 reduction was performed using a Gamry Interface 1000 electrochemical workstation. All electrochemical measurements were conducted in a three-electrode gas-sealed single-compartment glass cell, as shown in Figure S1. TiO2/MXene supported on CP (0.4 mg/cm2) was used as the working electrode while Ag/AgCl (0.1 M KPF6 in acetonitrile) and nickel foam (NF) (2 × 1 cm) electrodes were used as the reference and counter electrodes, respectively, and conducted in a 0.1 M KPF6 acetonitrile solution. The electrolyte solution was purged with CO2 for 30 min before measurement. Cyclic voltammogram (CV) scans were performed at 10 mV/s with IR compensation. Electrochemical impedance spectroscopy (EIS) analysis was performed at −1.9 V vs Ag/AgCl in the frequency range of 104–10−2 Hz. Controlled potential coulometry was performed at −1.9 V for 30 min. Gas products produced from electrolysis were measured by an SRI 8610C (SRI Instruments, Torrance, CA, USA) gas chromatograph equipped with a thermal conductivity detector (TCD) and a flame ionization detector (FID), using argon as the carrier gas. Faradaic efficiency (FE) for CO was calculated by
ε = e × F × M Q × 100 % ,
where ε is Faradaic efficiency, F = 96485   C / m o l is Faraday’s constant, e = 2 is the number of transferred electrons for CO2RR, M is the amount of CO (mol) and is obtained from the corresponding peak area of the FID curve from GC experiments, and Q is the charge obtained from the reaction.

4. Conclusions

In this work, we demonstrated the catalytic ability of novel TiO2/MXene nanocomposite electrodes for an efficient electrochemical CO2RR. The nanocomposite TiO2/MXene electrodes were easily fabricated using cost-effective solvothermal and drop-cast techniques. Electrochemical characterization of the electrode through cyclic voltammetry showed that it behaved as a capacitor in the absence of CO2 but a cathodic current in its presence indicated a reduction reaction at the electrode. The electrochemical performance of the CO2RR to produce CO showed an excellent FE of 99.41% at −1.9 V (vs. Ag/AgCl), by far outperforming the individual components. A DFT-level analysis of the CO2RR on TiO2/MXene electrodes revealed the likely mechanism producing CO and a carbonate anion through ∗CO, ∗O, and activated CO22− intermediates, as well as two less likely pathways which resulted in an oxalate or two CO molecules with one O2. Subsequent Bader charge analysis showed that TiO2/MXene behaves as a catalyst in the CO2RR compared to its TiO2 or MXene components alone. Additional work is still needed to confirm the proposed reaction mechanism by quantifying reaction products through mass-spectrometry and monitoring intermediate formation through sum-frequency generation spectroscopy. In sum, easy-to-produce TiO2/MXene nanocomposite electrodes may hold great utility in the harvesting and use of atmospheric CO2 through an efficient CO2RR to combat the pressing global climate and energy crises.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30194025/s1, Figure S1: Schematic of the CO2 reduction reaction electro-reduction cell; Figure S2. High-resolution XPS spectrum of F1s for TiO2/MXene; Figure S3. High-resolution XPS spectrum of F1s for TiO2/MXene; Figure S4. EDX result of TiO2/MXene; Figure S5. Faradaic efficiencies (FE) for CO by TiO2/MXene when electrolyte is saturated with CO2 (red) or N2 (green); Figure S6. Faradaic efficiencies (FE) for CO production by TiO2/MXene at different TiO2 growth times; Figure S7. The EIS curves of MXene, TiO2/MXene at solvothermal time of 1, 4, 12, and 36 h; Figure S8. Faradaic efficiencies (FE) for CO production by TiO2/MXene at different electrolysis time; Figure S9. (A) Surface energy of specified surfaces with miller indices of TiO2 and (B) the TiO2 nanoparticle constructed using Wulff construction method; Figure S10. Adsorption energy of TiO2 nanoparticle with various geometric orientations above MXene; Figure S11. Adsorption energy of two CO2 molecules on MXene at symmetrically distinct positions; Figure S12. Adsorption energy of a CO2 molecule on TiO2 at symmetrically distinct positions. Table S1. Comparison of electrocatalytic CO2RR activities of various nonprecious catalysts with those reported in the literature.

Author Contributions

Investigation, S.D., Z.S., A.S.K., F.S., T.Z., H.F., J.B.B. and Y.Q.; Resources, Y.R.; Writing–original draft, S.D., Z.S., A.S.K., F.S., H.F. and Y.R.; Writing–review & editing H.F., J.B.B., Z.-C.H.-F., H.W., Z.Z., M.S.M., R.L.O., Y.S. and Y.R.; Supervision, S.D., Z.S., A.S.K., F.S., Z.Z., M.S.M., R.L.O., Y.S. and Y.R.; Project administration, Z.Z. and Y.R.; Funding acquisition, Z.Z. and Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Reactive Chemical Systems Programs, U.S. Army Research Office (W911NF18020112). Funding was provided by the U.S. Army via the Surface Science Initiative Program (PE 0601102A Project VR9) at the US Army, Combat Capabilities Development Command, Chemical Biological Center. This work is also supported by the National Science Foundation under Grant No. [2045084].

Data Availability Statement

Data will be provided by the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  2. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  3. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  4. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual Cocatalysts in TiO2 Photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef] [PubMed]
  5. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  6. Peiris, S.; de Silva, H.B.; Ranasinghe, K.N.; Bandara, S.V.; Perera, I.R. Recent development and future prospects of TiO2 photocatalysis. J. Chin. Chem. Soc. 2021, 68, 738–769. [Google Scholar] [CrossRef]
  7. Lee, S.-Y.; Park, S.-J. TiO2 photocatalyst for water treatment applications. J. Ind. Eng. Chem. 2013, 19, 1761–1769. [Google Scholar] [CrossRef]
  8. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C-Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  9. Paz, Y. Application of TiO2 photocatalysis for air treatment: Patents’ overview. Appl. Catal. B-Environ. 2010, 99, 448–460. [Google Scholar] [CrossRef]
  10. Guo, Q.; Ma, Z.; Zhou, C.; Ren, Z.; Yang, X. Single Molecule Photocatalysis on TiO2 Surfaces. Chem. Rev. 2019, 119, 11020–11041. [Google Scholar] [CrossRef] [PubMed]
  11. Thompson, T.L.; Yates, J.T. Surface Science Studies of the Photoactivation of TiO2 New Photochemical Processes. Chem. Rev. 2006, 106, 4428–4453. [Google Scholar] [CrossRef]
  12. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  13. He, Y.; Tilocca, A.; Dulub, O.; Selloni, A.; Diebold, U. Local ordering and electronic signatures of submonolayer water on anatase TiO2(101). Nat. Mater. 2009, 8, 585–589. [Google Scholar] [CrossRef] [PubMed]
  14. Kershis, M.D.; Wilson, D.P.; White, M.G. Dynamics of acetone photooxidation on TiO2(110): State-resolved measurements of methyl photoproducts. J. Chem. Phys. 2013, 138, 204703. [Google Scholar] [CrossRef] [PubMed]
  15. Kondo, Y.; Yoshikawa, H.; Awaga, K.; Murayama, M.; Mori, T.; Sunada, K.; Bandow, S.; Iijima, S. Preparation, Photocatalytic Activities, and Dye-Sensitized Solar-Cell Performance of Submicron-Scale TiO2 Hollow Spheres. Langmuir 2008, 24, 547–550. [Google Scholar] [CrossRef]
  16. Chen, J.S.; Tan, Y.L.; Li, C.M.; Cheah, Y.L.; Luan, D.; Madhavi, S.; Boey, F.Y.C.; Archer, L.A.; Lou, X.W. Constructing hierarchical spheres from large ultrathin anatase TiO2 nanosheets with nearly 100% exposed (001) facets for fast reversible lithium storage. J. Am. Chem. Soc. 2010, 132, 6124–6130. [Google Scholar] [CrossRef]
  17. Zheng, Z.; Xie, W.; Lim, Z.S.; You, L.; Wang, J. CdS sensitized 3D hierarchical TiO2/ZnO heterostructure for efficient solar energy conversion. Sci. Rep. 2014, 4, 5721. [Google Scholar] [CrossRef]
  18. Li, D.; Xia, Y. Direct Fabrication of Composite and Ceramic Hollow Nanofibers by Electrospinning. Nano Lett. 2004, 4, 933–938. [Google Scholar] [CrossRef]
  19. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  20. Tan, S.S.; Zou, L.; Hu, E. Kinetic modelling for photosynthesis of hydrogen and methane through catalytic reduction of carbon dioxide with water vapour. Catal. Today 2008, 131, 125–129. [Google Scholar] [CrossRef]
  21. Kočí, K.; Obalová, L.; Šolcová, O. Kinetic study of photocatalytic reduction of CO2 over TIO2. Chem. Process Eng.-Inz. Chem. I Proces. 2010, 31, 395–407. [Google Scholar]
  22. Shkrob, I.A.; Zhu, Y.; Marin, T.W.; Abraham, D.P. Reduction of Carbonate Electrolytes and the Formation of Solid-Electrolyte Interface (SEI) in Lithium-Ion Batteries. 1. Spectroscopic Observations of Radical Intermediates Generated in One-Electron Reduction of Carbonates. J. Phys. Chem. C 2013, 117, 19255–19269. [Google Scholar] [CrossRef]
  23. Kreft, S.; Wei, D.; Junge, H.; Beller, M. Recent advances on TiO2-based photocatalytic CO2 reduction. EnergyChem 2020, 2, 100044. [Google Scholar] [CrossRef]
  24. Low, J.; Zhang, L.; Tong, T.; Shen, B.; Yu, J. TiO2/MXene Ti3C2 composite with excellent photocatalytic CO2 reduction activity. J. Catal. 2018, 361, 255–266. [Google Scholar] [CrossRef]
  25. VahidMohammadi, A.; Rosen, J.; Gogotsi, Y. The world of two-dimensional carbides and nitrides (MXenes). Science 2021, 372, eabf1581. [Google Scholar] [CrossRef] [PubMed]
  26. Gogotsi, Y.; Anasori, B. The Rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [PubMed]
  27. Gogotsi, Y. The Future of MXenes. Chem. Mater. 2023, 35, 8767–8770. [Google Scholar] [CrossRef]
  28. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed]
  29. Lim, K.R.G.; Shekhirev, M.; Wyatt, B.C.; Anasori, B.; Gogotsi, Y.; Seh, Z.W. Fundamentals of MXene synthesis. Nat. Synth. 2022, 1, 601–614. [Google Scholar] [CrossRef]
  30. Ahn, S.; Han, T.-H.; Maleski, K.; Song, J.; Kim, Y.-H.; Park, M.-H.; Zhou, H.; Yoo, S.; Gogotsi, Y.; Lee, T.-W. A 2D Titanium Carbide MXene Flexible Electrode for High-Efficiency Light-Emitting Diodes. Adv. Mater. 2020, 32, 2000919. [Google Scholar] [CrossRef]
  31. Zhang, Y.-Z.; Lee, K.H.; Anjum, D.H.; Sougrat, R.; Jiang, Q.; Kim, H.; Alshareef, H.N. MXenes stretch hydrogel sensor performance to new limits. Sci. Adv. 2018, 4, eaat0098. [Google Scholar] [CrossRef]
  32. Driscoll, N.; Richardson, A.G.; Maleski, K.; Anasori, B.; Adewole, O.; Lelyukh, P.; Escobedo, L.; Cullen, D.K.; Lucas, T.H.; Gogotsi, Y.; et al. Two-Dimensional Ti3C2 MXene for High-Resolution Neural Interfaces. ACS Nano 2018, 12, 10419–10429. [Google Scholar] [CrossRef]
  33. Ward, E.J.; Lacey, J.; Crua, C.; Dymond, M.K.; Maleski, K.; Hantanasirisakul, K.; Gogotsi, Y.; Sandeman, S. 2D Titanium Carbide (Ti3C2Tx) in Accommodating Intraocular Lens Design. Adv. Funct. Mater. 2020, 30, 2000841. [Google Scholar] [CrossRef]
  34. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D metal carbides and nitrides (MXenes) for energy storage. Nat. Rev. Mater. 2017, 2, 16098. [Google Scholar] [CrossRef]
  35. Li, X.; Huang, Z.; Shuck, C.E.; Liang, G.; Gogotsi, Y.; Zhi, C. MXene chemistry, electrochemistry and energy storage applications. Nat. Rev. Chem. 2022, 6, 389–404. [Google Scholar] [CrossRef]
  36. Nan, J.; Guo, X.; Xiao, J.; Li, X.; Chen, W.; Wu, W.; Liu, H.; Wang, Y.; Wu, M.; Wang, G. Nanoengineering of 2D MXene-Based Materials for Energy Storage Applications. Small 2021, 17, 1902085. [Google Scholar] [CrossRef]
  37. Xiong, D.; Li, X.; Bai, Z.; Lu, S. Recent Advances in Layered Ti3C2Tx MXene for Electrochemical Energy Storage. Small 2018, 14, 1703419. [Google Scholar] [CrossRef] [PubMed]
  38. Kajiyama, S.; Szabova, L.; Sodeyama, K.; Iinuma, H.; Morita, R.; Gotoh, K.; Tateyama, Y.; Okubo, M.; Yamada, A. Sodium-Ion Intercalation Mechanism in MXene Nanosheets. ACS Nano 2016, 10, 3334–3341. [Google Scholar] [CrossRef]
  39. Lukatskaya, M.R.; Mashtalir, O.; Ren, C.E.; Dall’Agnese, Y.; Rozier, P.; Taberna, P.L.; Naguib, M.; Simon, P.; Barsoum, M.W.; Gogotsi, Y. Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide. Science 2013, 341, 1502–1505. [Google Scholar] [CrossRef] [PubMed]
  40. Hu, M.; Li, Z.; Li, G.; Hu, T.; Zhang, C.; Wang, X. All-Solid-State Flexible Fiber-Based MXene Supercapacitors. Adv. Mater. Technol. 2017, 2, 1700143. [Google Scholar] [CrossRef]
  41. Zhu, J.; Tang, Y.; Yang, C.; Wang, F.; Cao, M. Composites of TiO2 Nanoparticles Deposited on Ti3C2 MXene Nanosheets with Enhanced Electrochemical Performance. J. Electrochem. Soc. 2016, 163, A785. [Google Scholar] [CrossRef]
  42. Wang, F.; Yang, C.; Duan, M.; Tang, Y.; Zhu, J. TiO2 nanoparticle modified organ-like Ti3C2 MXene nanocomposite encapsulating hemoglobin for a mediator-free biosensor with excellent performances. Biosens. Bioelectron. 2015, 74, 1022–1028. [Google Scholar] [CrossRef]
  43. Xiong, Z.-M.; O’Donovan, M.; Sun, L.; Choi, J.Y.; Ren, M.; Cao, K. Anti-Aging Potentials of Methylene Blue for Human Skin Longevity. Sci. Rep. 2017, 7, 2475. [Google Scholar] [CrossRef]
  44. Zhang, C.J.; Pinilla, S.; McEvoy, N.; Cullen, C.P.; Anasori, B.; Long, E.; Park, S.-H.; Seral-Ascaso, A.; Shmeliov, A.; Krishnan, D.; et al. Oxidation Stability of Colloidal Two-Dimensional Titanium Carbides (MXenes). Chem. Mater. 2017, 29, 4848–4856. [Google Scholar] [CrossRef]
  45. Shahzad, A.; Rasool, K.; Miran, W.; Nawaz, M.; Jang, J.; Mahmoud, K.A.; Lee, D.S. Mercuric ion capturing by recoverable titanium carbide magnetic nanocomposite. J. Hazard. Mater. 2018, 344, 811–818. [Google Scholar] [CrossRef] [PubMed]
  46. Shahzad, A.; Rasool, K.; Nawaz, M.; Miran, W.; Jang, J.; Moztahida, M.; Mahmoud, K.A.; Lee, D.S. Heterostructural TiO2/Ti3C2Tx (MXene) for photocatalytic degradation of antiepileptic drug carbamazepine. Chem. Eng. J. 2018, 349, 748–755. [Google Scholar] [CrossRef]
  47. Zhang, J.; Yang, L.; Wang, H.; Zhu, G.; Wen, H.; Feng, H.; Sun, X.; Guan, X.; Wen, J.; Yao, Y. In Situ Hydrothermal Growth of TiO2 Nanoparticles on a Conductive Ti3C2Tx MXene Nanosheet: A Synergistically Active Ti-Based Nanohybrid Electrocatalyst for Enhanced N2 Reduction to NH3 at Ambient Conditions. Inorg. Chem. 2019, 58, 5414–5418. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, Y.; Li, Y.-H.; Li, X.; Zhang, Q.; Yu, H.; Peng, X.; Peng, F. Regulating Electron–Hole Separation to Promote Photocatalytic H2 Evolution Activity of Nanoconfined Ru/MXene/TiO2 Catalysts. ACS Nano 2020, 14, 14181–14189. [Google Scholar] [CrossRef]
  49. Zhou, Y.; Chai, Y.; Li, X.; Wu, Z.; Lin, J.; Han, Y.; Li, L.; Qi, H.; Gu, Y.; Kang, L.; et al. Defect-Rich TiO2 In Situ Evolved from MXene for the Enhanced Oxidative Dehydrogenation of Ethane to Ethylene. ACS Catal. 2021, 11, 15223–15233. [Google Scholar] [CrossRef]
  50. Tahir, M.; Tahir, B. In-situ growth of TiO2 imbedded Ti3C2TA nanosheets to construct PCN/Ti3C2TA MXenes 2D/3D heterojunction for efficient solar driven photocatalytic CO2 reduction towards CO and CH4 production. J. Colloid Interface Sci. 2021, 591, 20–37. [Google Scholar] [CrossRef]
  51. 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]
  52. Tang, Q.; Xiong, P.; Wang, H.; Wu, Z. Boosted CO2 photoreduction performance on Ru-Ti3CN MXene-TiO2 photocatalyst synthesized by non-HF Lewis acidic etching method. J. Colloid Interface Sci. 2022, 619, 179–187. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, Y.; Wang, F.; Lei, S.; Wei, Y.; Zhao, D.; Gao, Y.; Ma, X.; Li, S.; Chang, S.; Wang, M.; et al. In situ grown two-dimensional TiO2/Ti3CN MXene heterojunction rich in Ti3+ species for highly efficient photoelectrocatalytic CO2 reduction. Chem. Eng. J. 2023, 452, 139392. [Google Scholar] [CrossRef]
  54. Khan, A.A.; Tahir, M.; Zakaria, Z.Y. Synergistic effect of anatase/rutile TiO2 with exfoliated Ti3C2TR MXene multilayers composite for enhanced CO2 photoreduction via dry and bi-reforming of methane under UV–visible light. J. Environ. Chem. Eng. 2021, 9, 105244. [Google Scholar] [CrossRef]
  55. Song, F.; Debow, S.; Zhang, T.; Qian, Y.; Huang-Fu, Z.-C.; Munns, K.; Schmidt, S.; Fisher, H.; Brown, J.B.; Su, Y.; et al. Interface Catalysts of Ni3Fe1 Layered Double Hydroxide and Titanium Carbide for High-Performance Water Oxidation in Alkaline and Natural Conditions. J. Phys. Chem. Lett. 2023, 14, 5692–5700. [Google Scholar] [CrossRef]
  56. Scarpelli, F.; Mastropietro, T.F.; Poerio, T.; Godbert, N. Mesoporous TiO2 Thin Films: State of the Art; InTech: Rijeka, Croatia, 2018. [Google Scholar]
  57. Gu, H.; Zhang, H.; Wang, X.; Li, Q.; Chang, S.; Huang, Y.; Gao, L.; Cui, Y.; Liu, R.; Dai, W.-L. Robust construction of CdSe nanorods@Ti3C2 MXene nanosheet for superior photocatalytic H2 evolution. Appl. Catal. B Environ. 2023, 328, 122537. [Google Scholar] [CrossRef]
  58. Zhang, Z.Y.; Chi, M.F.; Veith, G.M.; Zhang, P.F.; Lutterman, D.A.; Rosenthal, J.; Overbury, S.H.; Dai, S.; Zhu, H.Y. Rational Design of Bi Nanoparticles for Efficient Electrochemical CO2 Reduction: The Elucidation of Size and Surface Condition Effects. ACS Catal. 2016, 6, 6255–6264. [Google Scholar] [CrossRef]
  59. Allabour, A.; Coskun, H.; Zheng, X.L.; Kibria, M.G.; Strobel, M.; Hild, S.; Kehrer, M.; Stifter, D.; Sargent, E.H.; Stadler, P. Active Sulfur Sites in Semimetallic Titanium Disulfide Enable CO2 Electroreduction. ACS Catal. 2020, 10, 66–72. [Google Scholar] [CrossRef]
  60. Mendieta-Reyes, N.E.; Cheuquepán, W.; Rodes, A.; Gómez, R. Spectroelectrochemical Study of CO2 Reduction on TiO2 Electrodes in Acetonitrile. ACS Catal. 2020, 10, 103–113. [Google Scholar] [CrossRef]
  61. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  62. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  63. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  64. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865, Erratum in Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef] [PubMed]
  65. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  66. Kresse, G.; Joubert, D.P. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  67. Chávez-Rocha, R.; Mercado-Sánchez, I.; Vargas-Rodriguez, I.; Hernández-Lima, J.; Bazán-Jiménez, A.; Robles, J.; García-Revilla, M.A. Modeling Adsorption of CO2 in Rutile Metallic Oxide Surfaces: Implications in CO2 Catalysis. Molecules 2023, 28, 1776. [Google Scholar] [CrossRef]
  68. Liu, C.; Cundari, T.R.; Wilson, A.K. CO2 Reduction on Transition Metal (Fe, Co, Ni, and Cu) Surfaces: In Comparison with Homogeneous Catalysis. J. Phys. Chem. C 2012, 116, 5681–5688. [Google Scholar] [CrossRef]
  69. Chen, L.; Zhang, C.; Jiao, X. Recent Advances of In Situ Insights into CO2 Reduction toward Fuels. ChemCatChem 2025, 17, e202401388. [Google Scholar] [CrossRef]
  70. Pylnev, M.; Su, T.-S.; Wei, T.-C. Titania augmented with TiI4 as electron transporting layer for perovskite solar cells. Appl. Surf. Sci. 2021, 549, 149224. [Google Scholar] [CrossRef]
  71. Sang, X.; Xie, Y.; Lin, M.-W.; Alhabeb, M.; Van Aken, K.L.; Gogotsi, Y.; Kent, P.R.C.; Xiao, K.; Unocic, R.R. Atomic Defects in Monolayer Titanium Carbide (Ti3C2Tx) MXene. ACS Nano 2016, 10, 9193–9200. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of TiO2/MXene preparation procedure.
Figure 1. Schematic illustration of TiO2/MXene preparation procedure.
Molecules 30 04025 g001
Figure 2. (A) XRD spectra of TiO2/MXene, MXene, MAX, and TiO2 (top to bottom). High-resolution XPS spectra of (B) Ti 2p, (C) C 1s, and (D) O 1s for TiO2/MXene.
Figure 2. (A) XRD spectra of TiO2/MXene, MXene, MAX, and TiO2 (top to bottom). High-resolution XPS spectra of (B) Ti 2p, (C) C 1s, and (D) O 1s for TiO2/MXene.
Molecules 30 04025 g002
Figure 3. TEM (A) and HRTEM (B) images of TiO2/MXene nanocomposite. (C) Elemental mapping results of TiO2/MXene.
Figure 3. TEM (A) and HRTEM (B) images of TiO2/MXene nanocomposite. (C) Elemental mapping results of TiO2/MXene.
Molecules 30 04025 g003
Figure 4. Electrocatalytic performance of reduction of CO2 to CO. (A) Cyclic voltammograms (CVs) of TiO2/MXene and MXene in electrolyte saturated with N2 (black) and CO2 (red). Inset shows reference E0 value for (Fc/Fc+) standard; (B) controlled potential colometry (CPC) at potential of −1.9 V in electrolyte saturated with N2 (black) or CO2 (red); Faradaic efficiencies (FEs) of (C) TiO2/MXene at different potentials and (D) TiO2/MXene, MXene, and TiO2 at −1.9 V.
Figure 4. Electrocatalytic performance of reduction of CO2 to CO. (A) Cyclic voltammograms (CVs) of TiO2/MXene and MXene in electrolyte saturated with N2 (black) and CO2 (red). Inset shows reference E0 value for (Fc/Fc+) standard; (B) controlled potential colometry (CPC) at potential of −1.9 V in electrolyte saturated with N2 (black) or CO2 (red); Faradaic efficiencies (FEs) of (C) TiO2/MXene at different potentials and (D) TiO2/MXene, MXene, and TiO2 at −1.9 V.
Molecules 30 04025 g004
Figure 5. (A) The structure of the TiO2/MXene catalyst; (B) the adsorption energy of CO2 molecules above different titanium atoms of TiO2 nanoparticles on TiO2/MXene. (C) The energy profile of the reaction mechanism with MXene (brown), TiO2 (orange), and TiO2/MXene (green) as catalysts, with the end products of reaction paths noted above stage (III). The simulated intermediate structures of the TiO2/MXene system with the end products C O and C O 3 2 shown in the inset. (D) The charge density of adsorbed CO2 molecules on MXene, TiO2, and TiO2/MXene; (E) a side view of the Bader charge density illustration of adsorbed CO2 molecules on TiO2/MXene. Yellow and cyan denote electron accumulation and depletion.
Figure 5. (A) The structure of the TiO2/MXene catalyst; (B) the adsorption energy of CO2 molecules above different titanium atoms of TiO2 nanoparticles on TiO2/MXene. (C) The energy profile of the reaction mechanism with MXene (brown), TiO2 (orange), and TiO2/MXene (green) as catalysts, with the end products of reaction paths noted above stage (III). The simulated intermediate structures of the TiO2/MXene system with the end products C O and C O 3 2 shown in the inset. (D) The charge density of adsorbed CO2 molecules on MXene, TiO2, and TiO2/MXene; (E) a side view of the Bader charge density illustration of adsorbed CO2 molecules on TiO2/MXene. Yellow and cyan denote electron accumulation and depletion.
Molecules 30 04025 g005
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

Debow, S.; Shen, Z.; Kulathuvayal, A.S.; Song, F.; Zhang, T.; Fisher, H.; Brown, J.B.; Qian, Y.; Huang-Fu, Z.-C.; Wang, H.; et al. Interface Catalysts of In Situ-Grown TiO2/MXenes for High-Faraday-Efficiency CO2 Reduction. Molecules 2025, 30, 4025. https://doi.org/10.3390/molecules30194025

AMA Style

Debow S, Shen Z, Kulathuvayal AS, Song F, Zhang T, Fisher H, Brown JB, Qian Y, Huang-Fu Z-C, Wang H, et al. Interface Catalysts of In Situ-Grown TiO2/MXenes for High-Faraday-Efficiency CO2 Reduction. Molecules. 2025; 30(19):4025. https://doi.org/10.3390/molecules30194025

Chicago/Turabian Style

Debow, Shaun, Zichen Shen, Arjun Sathyan Kulathuvayal, Fuzhan Song, Tong Zhang, Haley Fisher, Jesse B. Brown, Yuqin Qian, Zhi-Chao Huang-Fu, Hui Wang, and et al. 2025. "Interface Catalysts of In Situ-Grown TiO2/MXenes for High-Faraday-Efficiency CO2 Reduction" Molecules 30, no. 19: 4025. https://doi.org/10.3390/molecules30194025

APA Style

Debow, S., Shen, Z., Kulathuvayal, A. S., Song, F., Zhang, T., Fisher, H., Brown, J. B., Qian, Y., Huang-Fu, Z.-C., Wang, H., Zander, Z., Mirotznik, M. S., Opila, R. L., Su, Y., & Rao, Y. (2025). Interface Catalysts of In Situ-Grown TiO2/MXenes for High-Faraday-Efficiency CO2 Reduction. Molecules, 30(19), 4025. https://doi.org/10.3390/molecules30194025

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop