Skip to main content
Ultrasonics Sonochemistry logoLink to Ultrasonics Sonochemistry
. 2023 Oct 5;100:106623. doi: 10.1016/j.ultsonch.2023.106623

WS2-intercalated Ti3C2Tx MXene/TiO2-stacked hybrid structure as an excellent sonophotocatalyst for tetracycline degradation and nitrogen fixation

Kugalur Shanmugam Ranjith a, Seyed Majid Ghoreishian b, Reddicherla Umapathi c, Ganji Seeta Rama Raju a, Hyun Uk Lee d, Yun Suk Huh c,, Young-Kyu Han a,
PMCID: PMC10585321  PMID: 37832252

Highlights

  • Ti3C2 MXene/TiO2-WS2 integrated heterostructure was prepared hydrothermally.

  • In-situ generated TiO2 on MXene-WS2 exhibited excellent sonophotocatalytic activity.

  • Excellent sonophotocatalytic NRR rate (526 μmol) is achieved by TiO2 interface on heterostructure.

  • Surface oxidized Ti-O strengthens the interface and promotes electron transfer path.

Keywords: MXene, Transition metal dichalcogenide, Sonophotocatalysis, N2 fixation, Interfacial contact

Abstract

Designing a heterostructure nanoscale catalytic site to facilitate N2 adsorption and photogenerated electron transfer would maximize the potential for photocatalytic activity and N2 reduction reactions. Herein, we have explored the interfacial TiO2 nanograins between the Ti3C2Tx MXene-WS2 heterostructure and addressed the beneficial active sites to expand the effective charge transfer rate and promote sonophotocatalytic N2 fixation. Benefiting from the interfacial contact and dual heterostructure interface maximizes the photogenerated carrier separation between WS2 and MXene/TiO2. The sonophotocatalytic activity of the MXene@TiO2/WS2 hybrid, which was assessed by examining the photoreduction of N2 with ultrasonic irradiation, was much higher than that of either sonocatalytic and photocatalytic activity because of the synergistic sonocatalytic effect under photoirradiation. The Schottky junction between the MXene and TiO2 on the hybrid MXene/TiO2-WS2 heterostructure resulted in the sonophotocatalytic performance through effective charge transfer, which is 1.47 and 1.24 times greater than MXene-WS2 for nitrogen fixation and pollutant degradation, respectively. Under the sonophotocatalytic process, the MXene/TiO2-WS2 heterostructure exhibits a decomposition efficiency of 98.9 % over tetracycline in 90 min, which is 5.46, 1.73, and 1.10 times greater than those of sonolysis, sonocatalysis, and photocatalysis, respectively. The production rate of NH3 on MXene/TiO2-WS2 reached 526 μmol g−1 h−1, which is 3.17, 3.61, and 1.47 times higher than that of MXene, WS2, and MXene-WS2, respectively. The hybridized structure of MXene-WS2 with interfacial surface oxidized TiO2 nanograins minimizes the band potential and improves photocarrier use efficiency, contributing directly to the remarkable catalytic performance towards N2 photo fixation under visible irradiation under ultrasonic irradiation. This report provides the strategic outcome for the mass carrier transfer rate and reveals a high conversion efficiency in the hybridized heterostructure.

1. Introduction

Global environmental and energy concerns have matured into severe issues in modern civilization that pose a grave danger to human life. Ammonia (NH3 production) is highly impactful for producing artificial fertilizers and sustainable carbon–neutral development progress [1], [2]. On an industrial scale, hydrogen and N2 gas are combined to generate NH3 by the traditional Haber–Bosch reaction process while being subjected to high temperatures and pressures in the presence of a catalyst [3]. This process would require a huge amount of fossil fuel resources, producing a high volume of greenhouse gases during the reaction that would impact global climate change [4]. To overcome this issue and explore a non-harmful route for NH3 production from the air, the N2 fixation of plants has inspired many researchers to develop the photocatalytic N2 fixation, which is a clean and non-harmful mild process and cost-effective process [5], [6]. Direct photo fixation of N2 at room temperature using a photocatalytic nitrogen reduction reaction (NRR) induced by renewable solar energy has the potential to synthesize NH3 in an environmentally friendly route [7]. In addition to the photocatalytic process, sonophotocatalytic technology has been studied extensively as a green technology for the degradation of pollutants and N2 fixation to promote the catalytic activity through the synergistic effect of the sonocatalytic and photocatalytic activity [8], [9]. Apart from the catalytic activity, the construction of innovative nanostructured materials leveraged the power of sonochemical approaches by generating heat and pressure in a short period of time [10], [11]. In a streamlined reaction environment with minimal energy usage, the ultrasonic process may be used to regulate the morphology and size of the nanostructures [12], [13], [14]. Sonochemical methods are widely used in the creation of nanostructures and in the removal of organic contaminants [15], [16]. An appealing strategy would be to take advantage of ultrasonication and visible, responsive catalytic activity that promotes the carrier density on the heterostructure interface by building a partially occupied band between the conduction band (CB) and valence band (VB) of a semiconductor catalyst.

Developing distinctive and efficient heterojunction photocatalysts for highly effective solar-to-chemical energy conversion is still challenging. The typical catalytic activity of the catalytic heterostructures is associated primarily with the band function and their effective carrier separation efficiency [17], [18], [19]. Among the various semiconductors, metal oxides, transition metal chalcogenides, and carbon derivatives are photocatalysts for degrading organic pollutants and N2 fixation activity [20], [21]. TiO2-based materials are the most stable catalytic network among the various semiconductor systems. On the other hand, their rapid carrier recombination and the poor visible response have limited their catalytic activity in the broad solar spectrum [22]. Inducing the doping functionality or tagging the other semiconductor as a heterostructure network has strategically improved the photocatalytic activity [23]. On the other hand, N2 reduction in an aqueous media is highly challenging through photocatalytic systems without sacrificial agents. The two major prerequisites are required for the effective carrier separation rate with high redox potential and reductive active sites. The ideal interface of the staggered band configuration of the heterostructure for effective charge carrier separation improvises selectively towards photoinduced N2 fixation [24], [25]. MXene, a metal carbide, is a contender in catalytic reactions because of its band configuration and high electrical conductivity, large surface area, and rich surface chemistry with tunable structural morphology [26]. Ti-based Ti3C2Tx MXene has attracted significant interest through its superior layered structure and considerable interest in photocatalysis, CO2 reduction, and water splitting through its band configuration [27], [28]. The three-layered Ti covering the two-layered C promoted the structural advantages of MXene for faster electron transportation [29]. On the other hand, MXene encounters several issues related to its easy aggregation through the weak van der Waals interrelation, oxidation, and surface defect formation in the aqueous solution phase, which limits their applicability for future development [30]. These issues have been addressed by conjugating the MXene with different carbon functionalities, polymers, and transition metal dichalcogenides to promote the functionality of MXene in vast applications [31], [32], [33]. Designing the MXene with the surface oxidized TiO2 nanograins influences the significant band functions with the Schottky junction. They promote the application towards photocatalysis in the form of Ti3C2/Ti3+-TiO2 where the contact interface of Ti3C2-TiO2 maximizes the photoinduced charge transfer and minimizes the carrier recombination [34]. In addition, MXene can partially oxidize to form the TiO2 on the MXene surface without titanium addition, leading to the Schottky barrier at the interface, and the heterostructure assembly would maximize the photoexcited carrier separation efficiency [35]. Inducing the defective sites and fluoride functionalities on the TiO2 surface promoted the activation of the catalytic surface for the N2 adsorption that maximizes the NH3 production rate as 206 μmol h−1 g−1, which is approximately 10 times larger than pure TiO2 [24]. Designing the TiO2 on the MXene surface as a hybrid structure (Ti3C2Tx/TiO2-400) exhibits the remarkable N2 photo fixation with the NH3 formation production rate as 422 μmol h−1 g−1 at room temperature because of the oxygen vacancy sites on the TiO2 phase that operate as active foci for N2 adsorption and activation.

Among the heterostructure interface of MXene, hexagonally packed transition metal atoms with two layered chalcogen atoms have significant advantages in energy and catalytic applications by offering distinctive optical, mechanical, and electrical behavior [36]. Controlling the morphology, highly active sites, and structural stability with high carrier mobility are the key drivers to maximize the functional advantage of the hybrid structure. Growing lamellar structures on the stacked MXene surface will fill up the space that inhibits restacking and exhibits the hierarchical structural assembly that promotes the structural advantages in photocatalytic and electrochemical applications [37]. Coupling the layered features of ZnIn2S4 in the MXene surface processes the porous scaffold hybrid configuration that reduces the electron transfer pathway and maximizes the photocatalytic activity [38], [39]. Engineering the stacked MXene-WS2 heterostructure with the surface promotes defect functionalities, strategically improvises the electron transfer rate at the heterostructure interface, and promotes the carrier migration path. Tagging WS2 over the MXene as a hierarchical structural assembly exhibited a significant improvement in electrochemical performances, which maximizes the specific capacitance of 1111F. g−1 @ 2 A g−1 with high energy and power density and evidence the electrocatalytic hydrogen evolution reaction at alkaline and acidic medium with ultralong stability of 24 h. The exhibition of the metallic 1 T phase and semiconducting 2H phase on WS2 with octahedral coordination has improved the carrier transfer rate, maximizing the photocatalytic efficiency through the improvised conductivity [40]. Enabling the heterostructure interface of WS2–TiO2 induces the type II band function that improves the carrier separation efficiency and the presence of defect coordinated that maximizes the photocatalytic conversion efficiency of N2 to NH3 at 1.39 mmol h−1 g−1 [41]. Hybridizing the multi-compound heterostructure feature to maximize seamless contact interface and minimize the charge transfer distance with less light shielding effectivity influences the carrier migration and promotes visible, responsive activity.

Inspired by previous reports, MXene-WS2 hierarchical heterostructures were constructed with oxygen-defected TiO2-based interfacial contact on the MXene surface to enrich the N2 concentration and promote the adsorption of N2 and catalytic active sites. The photocatalytic and sonocatalytic activity and selectivity were improved for the NRR-based NH3 synthesis by modulating the interfacial contact and the heterostructure assembly with the tunable surface interfacial states. The electronic configuration with the dual MXene/TiO2-WS2-based heterostructure interface maximizes the active area for the effective activation and adsorption of N2 owing to the interfacial contact of TiO2 through thermal oxidization. The thermally oxidized (300 °C) TiO2 on MXene-WS2 achieved the maximum sonophotocatalytic efficiency of 89.1 % under 60 min of visible irradiation, which was 2.47 and 1.24 times higher than MXene and MXene-WS2. Importantly, the superior sonophotocatalytic under visible light for the MXene/TiO2-WS2 evidences the NH3 production rate of 526 μmol·g−1·h−1, which is 2.55 and 1.44 times higher than sonolysis and photocatalysis. As an additional advantage, MXene/TiO2-WS2 also delivers high stability by maintaining > 94 % of its initial activity even after ten successive catalytic reactions. The local electron density of adsorbed N2 was tuned by tuning the local electronic states of oxygen vacancies by the thermally oxidized MXene surface, and the energy barrier for hydrogenation of N2 was minimized to promote the sonophotocatalytic N2 fixation. This study provides additional insights on the interfacial contact interface in designing the catalytic with multiple active sites for the photoresponsive NRR, and it reaffirms the electronic structural assembly to tune the catalytic activity.

2. Materials and methods

2.1. Chemicals

Ti3AlC2 powder was purchased from JILIN 11 Technology Co. LTD (China). Hydrofluoric acid (HF, 37 %), sodium tungstate dihydrate (Na2WO4·2H2O, 99 %), oxalic acid (C2H2O4, 98 %), thioacetamide (CH3CSNH2, 98 %), hydrochloric acid (HCl), benzoquinone (BQ, 98 %), silver nitrate (AgNO3, 99 %), tret-butanol (t-BuOH), disodium ethylenediaminetetraacetate (Na2-EDTA, 99 %), methanol (CH3OH), ethanol (C2H5OH), tetracycline (TC), sodium hydroxide (NaOH, 98 %), were obtained from Sigma–Aldrich. All the chemicals purchased were of analytical grade and used as received.

2.2. Synthesis of stacked MXene and oxidized MXene/TiO2 heterostructure

Stacked Ti3C2TX MXene was fabricated by acidic etching [42]. As a part of the etching strategy, 40 mL of a 37 % HF solution was taken in a Teflon vessel and stirred for 5 min, and 2 g of Ti3AlC2 MAX powder was added slowly to the acid solution to carve the Al from the MAX surface. The temperature of the solution was maintained at 30 °C for 72 h under constant stirring. The resulting solution was centrifuged, and the precipitate was washed with deionized (DI) water until the solution reached the desired pH of > 5. The resulting residue was collected and freeze-dried. Fig. 1a shows a schematic diagram of the process. MXene was oxidized by heat treating the as-prepared Ti3C2TX powder sample at 100, 200, 300, and 400 °C for 30 min in both ends closed tubular furnace to gain the TiO2 interface on the MXene surface. The samples were called MXene/TiO2(1 0 0), MXene/TiO2(2 0 0), MXene/TiO2, and MXene/TiO2(4 0 0). After thermal annealing, the surface of MXene was decomposed to TiO2, which minimized the structural quality of the stacked network and produced a Schottky junction between the MXene and TiO2.

Fig. 1.

Fig. 1

(a) Schematic synthesis route for fabricating the MXene/TiO2-WS2 hybrid heterostructure. (b) X-ray diffraction pattern and (c) FTIR spectra of MXene, MXene/TiO2, WS2, MXene-WS2, and MXene/TiO2-WS2 nanostructures.

2.3. Synthesis of WS2 nanoflakes

For the growth of WS2 nanoflakes, 5 mmol of thioacetamide and 2.5 mmol of Na2WO4·2H2O were dispersed in 30 mL of DI water and stirred homogeneously to attain a clear solution. Subsequently, 4 mmol of oxalic acid was used as a catalyst to control the pH of the solution as < 2, and the solution was stirred for 10 min to generate a homogeneous dispersion. The prepared solution was transferred into a Teflon-lined stainless-steel autoclave (50 mL) and heated for 12 h at 200 °C. After the reaction, the solution was naturally cooled to room temperature. The precipitate was collected, centrifuged with ethanol and DI water, and dried in vacuum annealing for 6 h at 60 °C.

2.4. Synthesis of MXene-WS2 and MXene/TiO2-WS2 hybrid structure

First, 40 mg of as-prepared MXene or MXene/TiO2(3 0 0) was dispersed homogeneously in the 30 mL of DI water and sonicated for 5 min at room temperature. The agitated suspension was added with the above 5 mmol of thioacetamide and 2.5 mmol Na2WO4·2H2O precursors and homogeneously stirred for 30 min. The mixed solution was placed into a Teflon-lined stainless-steel autoclave (50 mL) and heated for 12 h at 200 °C. The precipitate was recovered after cooling, centrifuged with ethanol and DI water, and dried in vacuum annealing for 6 h at 60 °C. Furthermore, the supporting information presents the detailed characterization, photocatalytic and sonophotocatalytic TC degradation, sonophotocatalytic N2 fixation measurements, and photoelectrochemical measurements. Based on the yield of WS2 synthesized through hydrothermal, the loading ratio of WS2 on MXene was varied, and their respective sonocatalytic properties were investigated. In addition, the WS2 was grown on different temperature-based oxidized MXene, and the catalytic activity of the composites was studied and compared.

2.5. Material characterization

X-ray diffraction (XRD, PANalytical X'Pert Pro multifunctional) was performed using Cu K irradiation (λ = 0.15406 nm) to characterize the structural properties of nanostructures. Fourier transform infrared (FTIR, JASCO FTIR 6600) spectroscopy was conducted to assess the surface functionalities of composites between 400 and 4000 cm−1. High-resolution scanning electron microscopy (HRSEM, S-4800, Hitachi), high-resolution transmission electron microscopy (FETEM, JEM-2100F, JEOL Japan), and selected area electron diffraction (SAED) were used to examine the morphologic and structural characterization of the produced composites. The specific surface areas and pore size distributions of composites were measured using a Tristar ASAP 2020 to measure the Brunauer–Emmett–Teller (BET) surface area and N2 adsorption at 77 K. A Thermo Scientific spectrometer with Al K (1486.6 eV) source were used for X-ray photoelectron spectroscopy (XPS). The UV-DRS measurements acquired with a UV–vis spectrophotometer (JASCO V-600) were used to estimate the optical adsorption and band gaps of the composites.

3. Results and discussion

As shown schematically in Fig. 1a, the MXene-WS2 hybrid heterostructure with an oxygen-deficient TiO2 interface was prepared using a three-step optimization process: (i) Initially, MXene was prepared by the selective etching of Al under the HF etching process. (ii) The MXene was oxidized thermally to induce the TiO2 nanograins on the layered surface with high oxygen defect states. (iii) Finally, the staked MXene surface as a hybrid heterostructure was decorated with WS2 nanoflakes through a hydrothermal process. Unless stated further, the discussion focuses on the characteristics of 300 °C of thermally oxidized TiO2 nanograins over MXene-WS2 heterostructure, which exhibits the best sono-photoreduction of NH3. Fig. 1b presents the corresponding XRD pattern of MXene and their respective compositions. The peaks associated with the (0 0 2), (0 0 6), and (0 0 8) crystal plains were evidence of the formation of MXene after the HF-based etching of MAX. With thermal oxidization, a broad peak at 25.2° was attributed to the anatase (1 0 1) plane of TiO2 nanograins on the MXene surface. Pure WS2 exhibits the (0 0 2), (0 0 4), (1 0 0), (1 0 3), and (1 0 5) lattice plains evidence of the hexagonal structure of WS2, which perfectly matches with JCPDS card no: 08–0237. Furthermore, the prepared MXene and oxidized MXene had the characteristic of (0 0 2) plane at 9.1°, which was shifted to 7.2° on the growing WS2 on the MXene surface, indicating the intercalation of MXene due to the growth of WS2 between the layers, which increased the d-spacing. Similarly, there was a mild shift in the WS2-based XRD peaks, evidencing the structural interaction of the heterostructure surface. The average crystalline size of the WS2 along the (0 0 2) plane for the MXene-WS2 and MXene/TiO2-WS2 samples was calculated as 18.2 and 17.6 nm, respectively, through the Scherrer formula [43], [44] showcases that there is no huge variation on the structural nature of the WS2 on the MXene surface through the inclusion of nanograin TiO2 features. The structural quality of the MXene and TiO2 was increased during the oxidization process, which resulted in intense peak signals. The FTIR spectra in Fig. 1c revealed the surface functional states of the heterostructure samples. The peak position between 3300 cm−1 to 3500 cm−1 was evidence of the stretching and bending vibrations of − OH functionalities, and the peak at approximately 1600 cm−1 revealed the bending vibrations of C–OH interaction on the MXene surface. Furthermore, the broadband at 482 cm−1, 576 cm−1, and 733 cm−1 were assigned to the Ti–C, Ti–O, and Ti–F interaction on the MXene samples, respectively. Although the surface was oxidized, the host layer of MXene was preserved with the Ti–C and Ti–F functionalities. The peaks at 1388 cm−1 and 1117 cm−1 were assigned to the O–H bending and C–F stretching vibrations, respectively. The vibrations of the –CH2, –C = C–, C–H, and C–O bonds were observed at 2920 cm−1, 1560 cm−1, 1458 cm−1, and 1120 cm−1, respectively. For WS2, the intense peaks from 623 cm−1 and 1080 cm−1 represent the W–S and S–S bonds, and the bands at 1396 cm−1 and 1632 cm−1 were attributed to the stretching vibrations of hydroxyl functionalities. The peaks at 3100 cm−1, 1200 cm−1, and 2100 cm−1 were assigned to the stretching and bending vibrations of W–S functionality. The specific surface area of the WS2, MXene, and MXene/TiO2-WS2 nanostructures was evaluated via the Brunauer–Emmett–Teller (BET) method, and the MXene/TiO2-WS2 heterostructure showed a higher surface area (5.84 m2/g) than MXene (2.43 m2/g) and WS2 (4.56 m2/g) nanostructures (Fig. S1). Intercalating of WS2 on the stacked MXene surface increases the surface area by creating high rough textured features on the stacked surface. The hierarchical assembly of nanoflakes on the stacked surface played an efficient role in improvising the surface area of the heterostructure assembly.

FESEM was conducted to examine the morphology and surface characteristics of the prepared heterostructure. Fig. 2 (a-c) shows FESEM images of the stacked MXene structures, which highlights the packed sheet assembly. This packed layered structure was formed due to the purposeful carving of Al from the MAX phase by the HF-based etching process. After thermally treating the MXene surface, small nano grain features appeared on the MXene surface due to TiO2 formation (Fig. 2 (d-f)). The grain assembly was disturbed uniformly throughout the stacked layer surface, indicating a TiO2-based heterostructure interface on the MXene surface. The hydrothermally prepared WS2 was evidence of a thin layered nanoflake-like assembly ((Fig. 2 g-i). In contrast, the growth of WS2 on the MXene surface resulted in the hierarchical hybrid heterostructure ((Fig. 2 j-l).

Fig. 2.

Fig. 2

FESEM images of (a-c) MXene, (d-f) MXene/TiO2, (g-i) WS2, and (j-l) MXene/TiO2-WS2 nanostructures in different magnifications.

During the hydrothermal process, the layered WS2 nanoflakes were grown epitaxially on the stacked MXene, resulting in a hierarchical sandwich heterostructure with the dual heterostructure interface between MXene, TiO2, and WS2. No significant morphological change in the heterostructure assembly was observed with and without the TiO2 formation on the MXene surface (Fig. S2). The loading density of WS2 on the MXene surface was increased as the WS2 concentration in the growth process was increased (Fig. S3). The high-magnified images of the heterostructures revealed the porous structure of the layered features with a nonuniform surface nature, which promotes the highly active sites and surface interaction during the catalytic reaction. The atomic level interaction and the structural assembly MXene, MXene/TiO2, WS2, and MXene/TiO2-WS2 were analyzed by TEM. Fig. 3 (a-c) shows TEM images of MXene at different magnifications. A smooth surface finish with the thin layered assembly was revealed from the high-magnified images. The derived SAED profile (Inset of Fig. 3b) indicated the polycrystalline nature of the MXene sheets. The MXene surface became rougher on thermally treating the samples at 300 °C with a dispersed nanograin assembly (Fig. 3 d-e) due to surface oxidization, which formed the TiO2 on the MXene surface. The HRTEM image of the MXene/TiO2 displayed two different sets of fringes with an interplanar spacing of 0.228 nm and 0.243 nm, representing the (1 0 3) plain of MXene and the (1 0 4) plain of TiO2 on the MXene/TiO2 sample. The co-existence of the TiO2 on the MXene surface was consistent with the XRD results. WS2 was exhibited as a thin sheet-like assembly, resulting in a lattice spacing of 0.621 nm representing the (0 0 2) crystallographic plains of WS2 (Fig. 3 g-i). Fig. 3(j-l) shows the hierarchical structure of the MXene-WS2 without an interlayer spacing between the MXene and WS2, evidence of the tightly coupled heterostructure interface. The HRTEM images of MXene/TiO2-WS2 heterostructure evidence the 0.262 nm lattice space corresponding to the (1 0 0) crystal plane of WS2. The coupling interface resulted in the lattice spacing of 0.228 nm representing the (1 0 3) crystal plane of MXene, showing the integrity of MXene/TiO2 and the strong contact interface of MXene/TiO2 and WS2. The EDAX profile of MXene/TiO2-WS2 shows the combined elemental peaks composed of 1.03, 49.16, 5.84, 0.33, 14.93, and 28.71 at.% of Ti, C, O, F, W, and S, respectively, which confirms the MXene/TiO2-WS2 nanocomposite (Fig. S4). The inset in Fig. 3k presents the SAED pattern of MXene/TiO2-WS2, showing the mixed crystal quality of polycrystalline WS2 and the crystalline network of MXene. Mapping analysis of MXene/TiO2-WS2 revealed the layered distribution of Ti, C, and O with the surface-loaded W and S elemental composition with a uniform distribution.

Fig. 3.

Fig. 3

Typical TEM and HRTEM images of (a-c) MXene; (d-f) MXene/TiO2; (g–i) WS2 and (j-l) MXene/TiO2-WS2 heterostructure. (m) EDAX mapping images of the MXene/TiO2-WS2 heterostructure. The inset shows the SAED pattern of the respective images.

The electronic states and elemental composition of MXene, WS2, MXene-WS2, and MXene/TiO2-WS2 were examined by X-ray photoelectron spectroscopy (XPS). Fig. S5 presents the survey spectrum of the MXene, WS2, MXene-WS2, and MXene/TiO2-WS2, showing C, Ti, O, W, and S. Fig. 4a shows the C 1 s spectra of MXene deconvoluted into multiple peaks at 281.7, 284.6, 285.4, 288.4, and 288.9 eV, which were assigned to the Ti–C, C–C, C–O, C Created by potrace 1.16, written by Peter Selinger 2001-2019 O, and C–F bonds, respectively [45]. The Ti 2p peaks of the pristine MXene peaks fitted with the Gaussian function at 454.9, 455.8, 456.9, and 458.8 eV were attributed to Ti–C (Ti+), Ti–X (Ti2+), Ti–X (Ti3+), and Ti–O, respectively [46]. Fig. 4c and 4d show the W 1 s and S 2p core level spectra of the pristine WS2 nanoflakes. The W 1 s spectra of WS2 nanoflakes show multiple Gaussian-fitted peaks attributed at 32.0, 34.1, and 38.6 eV, which corresponds to the binding energies of W 4f7/2, W 4f5/2, and W 5p3/2 that well matched with the W4+ species of WS2 [47]. On the other hand, an emerging peak at 35.4 eV was assigned to the W–O bond, revealing the significant oxidization of W on the nanoflakes [48]. The surface oxidization of W formed a part of unreacted W that was capped on the WS2 surface exposed to air. Surface oxidization will not influence the quality of the sample because the stochiometric of W:S is very close to the ideal value of 2. The doublet in S 2p spectra at 162.3 and 163.2 eV were attributed to S 2p3/2 and S 2p1/2 peaks of 1 T-WS2, representing S2- species. Fig. 4 (e-h) and 4(i-l) show the C1s, Ti 2p, W 4f, S 2p XPS spectra of the MXene-WS2 and MXene/TiO2-WS2 heterostructures. The oxidized sample has higher structural stability of the Ti–C phase after the hydrothermal process than the unoxidized sample. The durability of the Ti–C phase was increased by forming high crystalline TiO2 nanograins as a surface protector to minimize in-situ oxidization during the hydrothermal processes. The presence of interfacial TiO2 on the heterostructure surface is evidence of the dominant TiO2 vibration compared to the Ti–C and Ti3+ energy states. Unoxidized MXene-WS2 heterostructures also have the strong vibration of Ti–O, but the structural quality of MXene was poor because of the partial oxidization of MXene during the hydrothermal process. In addition, while growing the WS2 on the MXene surface, the rate of W oxidization was increased, which was observed from the exhibition of the intense peak at 35.4 eV on the W 4f peaks of heterostructure samples. The expanded W6+ states on the MXene-WS2 and MXene/TiO2-WS2 heterostructures indicated the interaction of the tungsten atom with distinct oxygen moieties from the environment [49]. In addition, there was no significant change in the S 2p spectra for the heterostructure samples, showing the good structural quality of WS2 on the MXene surface. From the observation, inducing thermal oxidization preserved the structural quality of the MXene surface during the hydrothermal process compared to the untreated MXene surface. The O 1 s spectra of the MXene/TiO2-WS2 heterostructure resulted in peaks at 529.6 and 531.3 eV, contributing to the Ti–O–Ti and Ti–O–H bonds, respectively (Fig. S6a). The peak at 684.8 eV was attributed to the Ti–F bonds, showing the F interaction on the MXene surface, as shown in Fig. S6b. The F functionality of the MXene remained even after decorating the WS2.

Fig. 4.

Fig. 4

High-resolution XPS profile: (a) C 1s and (b) Ti 2p of MXene; (c) W 4f and (d) S 2p of WS2; (e) C 1s, (f) Ti 2p, (g) W 4f, and (h) S 2p of MXene-WS2 heterostructure; (i) C 1s, (j) Ti 2p, (k) W 4f, and (l) S 2p of MXene/TiO2-WS2 heterostructure.

The optical absorption properties of MXene-based heterostructures were investigated through the UV–vis DRS spectrum to address the role of adsorption properties with the function of different heterostructure assemblies. Fig. S7 presents the visible light adsorption behavior on the heterostructure samples, and MXene/TiO2-WS2 showed high visible adsorption. All the samples exhibited high adsorption over the 280–750 nm range, with prominent plasmonic adsorption around 450 nm. The band edge adsorption moved toward the IR region in all the samples, which exhibited wide visible adsorption. The band gap of the materials was calculated using the Kubelka–Munk (K–M) plot. The estimated band gap of the WS2, MXene, MXene-WS2, and MXene/TiO2-WS2 heterostructures were 1.62, 1.59, 1.56, and 1.53 eV, respectively. The band gap decreased after inducing the TiO2 functionality as an interfacial layer on the heterostructure assembly.

3.1. Photocatalysis and sonophotocatalysis

The tetracycline (TC) removal efficiency was documented to address the sonophotocatalytic activity of the MXene/TiO2-WS2 hybrid heterostructure. Fig. 5a shows the UV–vis absorbance spectra of TC under different reaction conditions in 90 min. The effects of sonophotocatalytic properties were assessed by examining the sonolysis, adsorption, sonocatalysis, and photocatalysis properties of MXene/TiO2-WS2. Comparing the ultrasonic irradiation with and without MXene/TiO2-WS2, the adsorption peak intensity was reduced drastically, demonstrating the sonocatalytic activity of the heterostructure. In photo-irradiation, the rate of TC decomposition in solution was higher than the sonocatalytic activity, highlighting the promising effect of photo-responsive radical production on the catalytic surface. Combining the sonocatalytic and photocatalytic effects, the photoinduced reactive oxygen species oxidation and the pyrolysis caused by the cavitation effects from ultrasound promoted the degradation rate. MXene/TiO2-WS2 hybrid heterostructure showed 98.9 % TC decomposition efficiency in 90 min, which is significantly higher than that of adsorption (14.1 %), sonolysis (18.1 %), sonocatalysis (57.1 %), and photocatalysis (89.9 %) (Fig. 5b). The synergistic effects of the visible irradiation and ultrasonic contribution maximized the decomposition efficiency. In the dark, MXene/TiO2-WS2 exhibited a removal efficiency of 14.2 % due to the static interaction of surface charges and the presence of defect states that adsorb the pollutant ions. The higher separation efficiency of photogenerated carriers and reactive oxygen species produced during the light-induced sonophotocatalytic activity degraded the TC and finally mineralized organic pollutants into H2O and CO2. Owing to the high photoresponsive behavior, the MXene/TiO2-WS2 heterostructure catalyst resulted in high photocatalytic efficiency towards the TC compared to all other heterostructure assemblies (Fig. S8). Fig. 5c presents the sonophotocatalytic decomposition of TC over different catalysts. The sonophotocatalytic activity of the MXene/TiO2-WS2 resulted in a high removal efficiency, which was 2.48, 1.66, 1.98, and 1.24 times more elevated than that of MXene, MXene/TiO2, WS2, and MXene-WS2 after 60 min of irradiation. The sonophotocatalytic degradation efficiency of MXene, MXene/TiO2, WS2, MXene-WS2, and without catalyst were 36.2, 54.1, 45.3, 72.1, and 13.1 %, respectively, in 60 min of irradiation. From the observation, MXene/TiO2-WS2 showed high sonophotocatalytic degradation efficiency, and the sonophotocatalytic activity was observed in the following order: MXene/TiO2-WS2 > MXene-WS2 > WS2 > MXene/TiO2 > MXene. Tagging the WS2 nanoflakes as a heterostructure assembly on the MXene surface improved the sonophotocatalytic activity. The interfacial TiO2 nanograins between the MXene-WS2 interface minimized the conduction band position between the MXene and WS2. The TiO2 nanograins promoted the carrier separation efficiency between the heterostructure interface that helps maximize the catalytic efficiency. A degradation kinetic study of TC followed pseudo-first-order kinetics with correlation coefficient (R2) values > 0.99 (Fig. S9). The synergetic effect of MXene/TiO2-WS2 during the sonophotocatalytic process was addressed using the synergetic indices. The synergetic indices were calculated using eq. (1) from the estimated pseudo-first-order rate constant (kapp). The kapp values during sonolysis, adsorption, sonocatalysis, photocatalysis, and sonophotocatalysis for TC using MXene/TiO2-WS2 were 0.0019, 0.0011, 0.0078, 0.0171, and 0.0341 min−1, respectively (Fig. 5d). Degradation turnover (dTON) was proposed to quantify the catalytic degradation activity to compare catalysts regardless of pollutant concentration or catalyst quantity. The dTON for the control experiment was calculated using the equation below [50], [51].

dTON=Mi-[Mf]t×[Cat]

Fig. 5.

Fig. 5

UV–Vis spectra of TC with a catalyst under different reaction conditions in 90 min. (b) the degradation efficiency of TC with MXene/TiO2-WS2 catalyst by absorption, sonolysis, sonocatalysis, photocatalysis, and sonophotocatalysis. (c) Effects of different MXene compositions on the sonophotocatalytic degradations of TC. (d) Synergy factors for the sonophotocatalytic degradations of TC by MXene/TiO2-WS2. Sonophotocatalytic efficiency over (e) different WS2 loading densities and (f) different oxidization temperatures of MXene/TiO2 (100–400 °C). Operational conditions: [MXene/TiO2-WS2] = 1 g L−1, TC = 10 mg L−1, and neutral pH.

Mi and Mf are the pollutant concentrations before and after the sonocatalytic process, respectively; t is the reaction time (h); and Cat signifies the amount of catalyst (g). The dTON of MXene/TiO2-WS2 was estimated as 1.57 µmol h−1 gcat−1 for TC, which is significantly higher than that of MXene/TiO2 (0.95 µmol h−1 gcat−1) and MXene-WS2 (1.28 µmol h−1 gcat−1). Introducing the interfacial TiO2 nanograins at the heterostructure interface of MXene-WS2 significantly boosted the sonophotocatalytic activity of heterostructure through the effective photogenerated charge carrier mobility. In addition, with increasing WS2 loading density on the MXene surface, the sonophotocatalytic performance of MXene/TiO2-WS2 was enhanced initially and decreased (Fig. 5e). The optimized growth rate of MXene/TiO2-WS2 composition showed high decomposition efficiency and excessive WS2 on the MXene had a shielding effect that led to a decrease in the sonophotocatalytic performance of MXene/TiO2-WS2. The sonophotocatalytic efficiency was improved by increasing the oxidization temperature of MXene before WS2 growth (Fig. 5f). The oxidization temperature was varied from 100 to 400 °C. Increasing the oxidization temperature to ≤ 300 °C increased the degradation rate, possibly because of the influence of possible oxygen defect functionalities on the TiO2 nanograins that promote the catalytic activity by inducing the defect band level on the heterostructure network. Although increasing the oxidization temperature to 400 °C, possible defect density was reduced compared to the low-temperature oxidization that promotes the carrier recombination rate that would minimize the catalytic efficiency. The effectiveness of the MXene/TiO2-WS2 composition over the organic pollutant degradation is compared and summarized in Table S1. The Ti-base MAX and layered MXene-based nanostructures were evidence of the significant sonocatalytic activity over the degradation of pharmaceutical pollutants [52], [53]. The production rate of reactive oxygen species and boosted H2O2 production has maximized the degradation rate with respect to time. In contrast, the MXene/TiO2-WS2 heterostructure sped up the reaction and promoted the degradation efficiency of 98.9 % in 90 min.

The sonophotocatalytic fixation of N2 to NH3 was studied under ambient conditions for 130 min under visible light irradiation. Fig. 6a presents the sonophotocatalytic synthesis of nitrogen fixation over the different samples. Among the all-other catalysts, the MXene/TiO2-WS2 heterostructure exhibited a higher N2 fixation ability than the WS2, MXene, and MXene/TiO2. The rate of NH3 production increased through the influence of the TiO2 interfacial layer, and the 300 °C thermal treatment reached a high NH3 production rate. This may be because of the high photo-responsive nature and promotive carrier separation efficiency. The rate of NH3 formation over the MXene, MXene/TiO2, WS2, MXene-WS2, and MXene/TiO2-WS2 catalysts were 169, 256, 146, 356, and 526 μmol h−1 g−1, respectively. Fig. 6b compares the effects of the ultrasonic condition, photocatalytic activity, and their dual effect with different operation conditions with MXene/TiO2-WS2 as a catalyst. The sonophotocatalytic nitrogen fixation (Ultra + light + cat + N2) efficiency of MXene/TiO2-WS2 (526 μmol h−1 g−1) is higher than the photocatalysis (Light + cat + N2) (366 μmol h−1 g−1) and sonocatalysis (Ultra + cat + N2) (206 μmol h−1 g−1), which is 1.44 and 2.55 times higher than that of photocatalysis and sonocatalysis, respectively. While comparing with the sonophotocatalysis nitrogen fixation, the photocatalytic active activity in the presence of stirring did not show any significant N2 fixation activity compared to photocatalysis, which was attributed to the hot spot effect through ultrasonic irradiation. The sonophotocatalytic experiments were performed under an Ar atmosphere instead of N2 to assess the influential role of N2 in the catalytic reaction. The results revealed suppressed NH3 formation, showing that N2 is the nitrogen source for NH3 production. Under photo-irradiation, the photogenerated electrons are the main active species for N2 fixation. Moreover, the photogenerated electrons and hole were well separated and effectively contributed to the reduction process, which maximizes the catalytic activity due to the high visible adsorption and ternary band function. Fig. 6d shows the impact of ultrasonic power on sonophotocatalytic N2 fixing. The enhanced cavitation effect increasing the NH3 formation rate was due to the extended sonication power from 90 W to 180 W. Increasing the sonication power to 210 W did not increase NH3 formation significantly compared to 180 W. This may be because increasing the sonication power tends to increase the generation of bubbles too large to continue the cavitation [54], which may minimize the N2 production rate. In addition, the sonophotocatalytic performance, catalytic usability, and stability are the concern parameters for the catalyst for the practical application. The rate of NH3 production has been monitored as 498 μmol h−1 g−1 in 120 min in the fourth cycle, indicating a stable catalyst (Fig. 6e). The loss in the catalytic activity may be due to some mass loss of catalyst during the reusable process. In addition, XRD showed that the structural nature of the MXene/TiO2-WS2 was preserved, even after four reusable cycles, and the morphology of the reused catalyst did not show any significant change (Fig. 6f). The sonophotocatalytic N2 fixation properties of MXene/TiO2-WS2 heterostructure were compared with the recently reported heterostructure catalysts (Table S2). The production rate of NH3 is higher for the ternary MXene/TiO2-WS2 heterostructure as compared with most of the heterostructure catalysts. The higher efficiency of the catalytic system can be attributed to the highly photo-responsive nature, structure of the heterostructure interface, and decrease in the resistance of carrier transfer by creating the MXene-WS2 heterostructure with TiO2 nanograin interface. The photoluminescence (PL) properties of the MXene and their heterostructures were studied to investigate the charge transfer effect in the heterostructure and recombination effect. Fig. 7a presents the PL spectra of MXene heterostructure under the excited wavelength of 375 nm. Compared to the MXene and WS2 samples, the heterostructure samples had a low PL intensity, suggesting a reduced recombination rate for electron and hole pairs produced by photosynthesis. Under light irradiation, the photoinduced electrons were separated from the CB of WS2 to the MXene via the CB of TiO2, which promotes separation efficiency by minimizing the recombination rate. The possible charge transfer mechanism was assessed by obtaining the XPS valance band spectra and Mott–Schottky curve to study the band details of WS2 and MXene-WS2 and MXene/TiO2-WS2.

Fig. 6.

Fig. 6

(a)Yield of NH3 concentration over different catalysts during the sonophotocatalytic process respective to time. (b) NH3 concentration and (c) NH3 production rate under different conditions with the MXene/TiO2-WS2 catalyst. (d) sonication power on the formation of NH3, (e) cycling experiment on MXene/TiO2-WS2 catalyst. (f) XRD and SEM image of the MXene/TiO2-WS2 catalyst after four reusable catalytic cycles.

Fig. 7.

Fig. 7

(a) Photoluminescence spectra of the as-prepared MXene-based heterostructures. (b) XPS valance band spectra of WS2, MXene-WS2, and MXene-TiO2-WS2. (c) Mott Schottky plot of the WS2, MXene/TiO2, MXene-WS2, and MXene/TiO2-WS2. (d) Transition photocurrent response of WS2, MXene-WS2, and MXene/TiO2-WS2 under visible irradiation using three-electrode measurements. (e) Electrochemical impedance spectrometry (EIS) spectra of the MXene-based heterostructure. (f) Sonophotocatalytic nitrogen fixation of MXene/TiO2-WS2 in the presence of alcohol hole scavengers (methanol, ethanol, and tert-butanol) and electron scavengers (AgNO3).

From the XPS VB (Fig. 7b) spectra, the VB of the WS2, MXene-TiO2, and MXene/TiO2-WS2 were estimated to be 0.60, 1.94, and 0.86 eV, respectively. The Mott–Schottky curve (Fig. 7c) was analyzed over the heterostructure catalyst to calibrate the band structure. The flat band potential WS2, MXene-WS2, and MXene/TiO2-WS2 were estimated to be − 0.63, −0.49, and − 0.54 eV (vs. Ag/AgCl) based on the linear potential intersects at Cs2 = 0. The positive slope of the WS2, MXene-WS2, and MXene/TiO2-WS2 confirmed the n-type behavior of the nanostructures. The transition photocurrent of the as-prepared samples was analyzed under visible irradiation with 50 sec on/off cycles. Compared to WS2, the MXene-WS2 and MXene/TiO2-WS2 heterostructures have displayed higher photocurrent density because of the improved charge separation efficiency of the photoinduced charge carrier. The photocurrent density of MXene/TiO2-WS2 was more than 10 times higher than the WS2 nanoflakes, addressing the significant heterostructure interface and carrier migration efficiency. Furthermore, the flat band potential of the n-type materials was equal to the Fermi level, where CB was very close to the Fermi level. Moreover, the VB position was calculated from the XPS VB spectra. The calculated band gap from the UV–Vis DRS results was correlated with the VB position, and the CB was calculated. As shown in Fig. 7d, the photocurrent density of MXene-WS2 and MXene/TiO2-WS2 samples were spiked immediately under visible irradiation and dropped under dark conditions. The photocurrent density of MXene/TiO2-WS2 was approximately two times higher than that of MXene-WS2. The charge transfer resistance between the electrolyte and the catalyst was studied by EIS. Fig. 7e shows the Nyquist plot attained from the catalyst without visible light irradiation. The arc radius of MXene/TiO2-WS2 was smaller than the other catalysts, highlighting the low charge transfer resistance. The electrolyte resistance (Rs) and charge transfer resistance Rct were attributed to the internal charge transfer resistance of the catalytic nanostructures. Fig. 7e shows the Nyquist plot attained from the catalyst without visible light irradiation. The arc radius of MXene/TiO2-WS2 was smaller than the other catalysts, highlighting the low charge transfer resistance. The electrolyte resistance (Rs) and charge transfer resistance Rct were attributed to the internal charge transfer resistance of the catalytic nanostructures. Fig. 7e shows that MXene/TiO2-WS2 has the Rct value of 236 Ω, which is smaller than the MXene-WS2 and MXene/TiO2 nanostructures. The result shows that tagging the WS2 on the MXene surface minimizes the charge transfer resistance. The in-situ generation of TiO2 facilitates the charge transfer and promotes the photogenerated carrier across the electrode and electrolyte interface. Fig. 7f shows the scavenger experiment to understand the reaction mechanism of the MXene/TiO2-WS2 sonophotocatalyst during the photo fixation process of N2. AgNO3 and alcohol (methanol, tert-butanol, and ethanol) were trapping agents for e and h+, respectively. The N2 fixation process was suppressed by adding AgNO3, suggesting that the photogenerated electrons are the promising active species in the photo fixation of N2. The possible sonophotocatalytic N2 fixation process was explained as follows:

H2O + ultrasound → •OH + •H
MXene/TiO2-WS2 + photo + sono energy → e + h+
e + O2 →.O2
h+ + OH →.OH
2H2O + 4 h+ → O2 + 4H+
N2 + 6H+ + 6 e → 2NH3
NH3 + H2O → NH3·H2O ↔ NH4+ + OH

Fig. 8 shows the proposed possible sonophotocatalytic N2 fixation over MXene/TiO2-WS2. The synergistic function of the sonocatalysis and photocatalysis promotes the sonophotocatalytic functionality of the MXene/TiO2-WS2 heterostructure. The band position of WS2 was calculated using the XPS valance band position and Mott–Schottky plot. The valance band positions of WS2 and MXene/TiO2 were 0.60 eV (0.51 vs. NHE) and 1.94 eV (1.85 vs. NHE), respectively. According to the UV DRS results, the calculated CB position was − 1.11 eV vs. NHE for WS2. TiO2 formed over the MXene would form a Schottky junction, and the work function of MXene was greater than that of WS2, resulting in the electron in WS2 flowing through the metallic MXene, leading to the effective carrier separation. In addition, the work function of MXene would vary depending on the surface functionality where the hydrothermal growth of WS2 could induce the − O or − OH functionalities [55]. The strong interfacial contract of TiO2 and MXene established the Schottky contact, and the surface-loaded WS2 resulted in the type II band function between the WS2 and in-situ-generated TiO2 nanograins. To understand the insight of MXene/TiO2-WS2 heterostructure, the VB and CB position of WS2 is calculated as 0.6 eV and −1.02 eV from the XPS VB spectra and their respective bandgap calculated from UV-DRS spectra. The VB and CB positions of TiO2 were determined as 2.45 V and − 0.63 V vs. NHE, according to a previous report [56]. Under photoirradiation, electrons from WS2 were excited from the VB to CB. Because the CB of WS2 is more negative than the TiO2 nanograin, photogenerated electrons migrated from the CB of WS2 to TiO2, and transferred to the MXene, which is more positive than the TiO2 and WS2 when the contact among the TiO2, WS2, and MXene is so tight [41].

Fig. 8.

Fig. 8

Sonophotocatalytic nitrogen fixation mechanism of the MXene/TiO2-WS2 heterostructure.

The presence of oxygen vacancy states on the TiO2 will trap the electrons in the CB and inject them to the empty antibonding orbitals (π*) of the N2 molecule, which promotes the formation of NH3. Hence, the highly crystalline TiO2 on the MXene surface has a co-catalytic effect that could extract the photogenerated electrons rapidly for the fixation of N2 to boost NH3 production. Cavitation bubble formation caused by acoustic cavitation by the sonication can generate hot spots that can pyrolyze the water to generate •OH. The influence of ultrasonication emitted the sonoluminescence that excited the e that further reduce the N2 to NH3. In the absence of catalysis, the sonolysis can explain the role of the ultrasonic wave that influences the cavitation effect for NH3 production. The oxidization of MXene induced TiO2 functionality on the surface with high oxygen-based defect states, causing extended visible light adsorption. When forming the heterostructure interface of WS2 on the MXene/TiO2 surface, the electrons are photoexcited to the CB of WS2 and CB of defect TiO2 nanograins under light irradiation. With the type II band function, the photogenerated electrons were moved towards the CB of TiO2 and transferred to the MXene surface through the Schottky junction interface. The hole was well separated in the VB of WS2, which minimized carrier recombination. Because the CB position of MXene was more negative than N2/NH3 redox potential (−0.092 V vs. NHE), which effectively reduces the N2 to NH3. The rate of photoinduced electron separation was influenced by photogenerated holes collected on the VB of WS2 and TiO2, where the VB holes oxidized the H2O to provide H+ for NH3 production. Sonophotocatalytic generated electrons and holes might react with surface adsorbed OH and O2 to produce •OH and •O2 radicals that decompose the organic contaminants into H2O and CO2. The as-prepared MXene/TiO2-WS2 heterostructure shows enhanced N2 fixation performances through the influence of sonication in the traditional photocatalytic system.

4. Conclusion

In summary, a highly active hybrid hierarchical MXene/TiO2-WS2 heterostructure sonophotocatalyst was designed and fabricated through the thermal treatment of MXene followed by the hydrothermal process. This strategy endows the in-situ generation of TiO2 nanograins on the stacked MXene surface that decorated by the hierarchical growth of WS2 nanoflakes. During photo fixation of N2, MXene/TiO2-WS2 achieved the maximum NH3 production rate of 526 μmol·g−1·h−1 under a visible, responsive sonophotocatalytic reaction, which was 2.06 and 1.47 times higher than MXene/TiO2 and MXene-WS2. Further, the sonophotocatalytic decomposition efficiency of TC was 2.19 and 1.32 times higher than the sonocatalytic and photocatalytic degradation rates. The defect-enriched TiO2-based interfacial layer on the MXene enriched the electronic configuration with the Schottky junction, and the dual MXene/TiO2-WS2 heterostructure interface promoted the effective activation/adsorption of N2. It was found that, MXene/TiO2-WS2 heterostructure present a wide adsorption range with favorable band potential for effective charge separation rate that increases the radical production under light irradiation. Furthermore, the rate of NH3 production by the MXene/TiO2-WS2 heterostructure retained 94 % after four cycles, indicating the durability and reusability of the sonophotocatalytic process. The hybridized MXene/TiO2-WS2 heterostructure improves photocarrier usage and leads to superior catalytic performance for N2 photo fixation and TC decomposition. The present finding focuses on developing new active catalytic through the ratio design of catalytic sites to increase the adsorption of N2 concentration on the catalytic surface and effective carrier separation efficiency to achieve high N2 conversion efficiency.

CRediT authorship contribution statement

Kugalur Shanmugam Ranjith: Investigation, Formal analysis, Conceptualization, Writing – original draft. Seyed Majid Ghoreishian: Software, Formal analysis. Reddicherla Umapathi: Formal analysis. Ganji Seeta Rama Raju: Formal analysis. Hyun Uk Lee: Formal analysis. Yun Suk Huh: Supervision, Writing – review & editing. Young-Kyu Han: Supervision, Writing – review & editing.

Declaration of Competing Interest

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

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean Government (MSIT) (RS-2023-00243617, 2022M3J7A1062940, 2022R1C1C1010601, 2022R1A2C2008968).

Footnotes

Appendix A

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ultsonch.2023.106623.

Contributor Information

Yun Suk Huh, Email: yunsuk.huh@inha.ac.kr.

Young-Kyu Han, Email: ykenergy@dongguk.edu.

Appendix A. Supplementary data

The following are the Supplementary data to this article:

Supplementary data 1
mmc1.docx (1.9MB, docx)

Data availability

Data will be made available on request.

References

  • 1.Kim K., Zagalskaya A., Ng J.L., Hong J., Alexandrov V., Pham T.A., Su X. Coupling nitrate capture with ammonia production through bifunctional redox-electrodes. Nat. Commun. 2023;14:823. doi: 10.1038/s41467-023-36318-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ghoreishian S.M., Shariati K., Huh Y.S., Lauterbach J. Recent advances in ammonia synthesis over ruthenium single-atom-embedded catalysts: A focused review. Chem. Eng. J. 2023;467 doi: 10.1016/j.cej.2023.143533. [DOI] [Google Scholar]
  • 3.Kim J.H., Dai T.Y., Yang M., Seo J.M., Lee J.S., Kweon D.H., Lang X.-Y., Ihm K., Shin T.J., Han G.-F., Jiang Q., Baek J.-B. Achieving volatile potassium promoted ammonia synthesis via mechanochemistry. Nat. Commun. 2023;14:2319. doi: 10.1038/s41467-023-38050-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kyriakou V., Garagounis I., Vourros A., Vasileiou E., Stoukides M. An Electrochemical Haber-Bosch Process. Joule. 2020;4:142–158. doi: 10.1016/j.joule.2019.10.006. [DOI] [Google Scholar]
  • 5.Liu Z., Fan S., Li X., Niu Z., Wang J., Bai C., Duan J., Tadé M.O., Liu S. Synergistic effect of single-atom Cu and hierarchical polyhedron-like Ta3N5/CdIn2S4 S-scheme heterojunction for boosting photocatalytic NH3 synthesis. Appl. Catal. B. 2023;327 doi: 10.1016/j.apcatb.2023.122416. [DOI] [Google Scholar]
  • 6.Morawski A.W., Cmielewska K., Ekiert E., Kusiak Nejman E., Pełech I., Staciwa P., Sibera D., Wanag A., Kapica-Kozar J., Gano M., Lendzion-Bieluń Z., Narkiewicz U. Effective green ammonia synthesis from gaseous nitrogen and CO2 saturated-water vapour utilizing a novel photocatalytic reactor. Chem. Eng. J. 2022;446 doi: 10.1016/j.cej.2022.137030. [DOI] [Google Scholar]
  • 7.Vu M.H., Quach T.A., Do T.O. The construction of Ru-doped In2O3 hollow peanut-like structure for an enhanced photocatalytic nitrogen reduction under solar light irradiation. Sustain. Energy Fuels. 2021;5:2528–2536. doi: 10.1039/D1SE00257K. [DOI] [Google Scholar]
  • 8.Maimaitizi H., Abulizi A., Zhang T., Okitsu K., Zhu J.-J. Facile photo-ultrasonic assisted synthesis of flower-like Pt/N-MoS2 microsphere as an efficient sonophotocatalyst for nitrogen fixation. Ultrason. Sonochem. 2020;63 doi: 10.1016/j.ultsonch.2019.104956. [DOI] [PubMed] [Google Scholar]
  • 9.Ding Z., Sun M., Liu W., Sun W., Meng X., Zheng Y. Ultrasonically synthesized N-TiO2/Ti3C2 composites: Enhancing sonophotocatalytic activity for pollutant degradation and nitrogen fixation. Sep. Purif. Technol. 2021;276 doi: 10.1016/j.seppur.2021.119287. [DOI] [Google Scholar]
  • 10.Ajabshir S.Z., Asil S.A.H., Niasari M.S. Recyclable magnetic ZnCo2O4-based ceramic nanostructure materials fabricated by simple sonochemical route for effective sunlight-driven photocatalytic degradation of organic pollution. Ceram. Int. 2021;47:8959–8972. doi: 10.1016/j.ceramint.2020.12.018. [DOI] [Google Scholar]
  • 11.Mahdavi K., Ajabshir S.Z., Yousif Q.A., Niasari M.S. Enhanced photocatalytic degradation of toxic contaminants using Dy2O3-SiO2 ceramic nanostructured materials fabricated by a new, simple and rapid sonochemical approach. Ultrason. Sonochem. 2022;82 doi: 10.1016/j.ultsonch.2021.105892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ajabshir S.Z., Emsaki M., Hosseinzadeh G. Innovative construction of a novel lanthanide cerate nanostructured photocatalyst for efficient treatment of contaminated water under sunlight. J. Colloid Interface Sci. 2022;619:1–13. doi: 10.1016/j.jcis.2022.03.112. [DOI] [PubMed] [Google Scholar]
  • 13.Zonarsaghar A., Kamazani M.M., Ajabshir S.Z. Sonochemical synthesis of CeVO4 nanoparticles for electrochemical hydrogen storage. Int. J. Hydrogen Energy. 2022;47:5403–5417. doi: 10.1016/j.ijhydene.2021.11.183. [DOI] [Google Scholar]
  • 14.Ajabshir S.Z., Baladi M., Niasari M.S. Sono-synthesis of MnWO4 ceramic nanomaterials as highly efficient photocatalysts for the decomposition of toxic pollutants. Ceram. Int. 2021;47:30178–30187. doi: 10.1016/j.ceramint.2021.07.197. [DOI] [Google Scholar]
  • 15.Rad S.S., Khataee A., Oskoui S.A., Rad T.S., Zarei M., Orooji Y., Gengec E., Kobya M. Carbonaceous CoCr LDH nanocomposite as a light-responsive sonocatalyst for treatment of a plasticizer-containing water. Ultrason. Sonochem. 2023;98 doi: 10.1016/j.ultsonch.2023.106485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Xu L., Liu N.P., An H.L., Ju W.T., Liu B., Wang X.F., Wang X. Preparation of Ag3PO4/CoWO4 S-scheme heterojunction and study on sonocatalytic degradation of tetracycline. Ultrason. Sonochem. 2022;89 doi: 10.1016/j.ultsonch.2022.106147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Gu H., Zhang H., Wang X., Li Q., Chang S., Huang Y., Gao L., Cui Y., Liu R., Dai W.-L. Xinglin Wang, Qin Li, Shengyuan Chang, Yamei Huang, Linlin Gao, Yuanyuan Cui, Renwei Liu, Wei-Lin Dai, Robust construction of CdSe nanorods@Ti3C2 MXene nanosheet for superior photocatalytic H2 evolution. Appl. Catal. B. 2023;328:122537. doi: 10.1016/j.apcatb.2023.122537. [DOI] [Google Scholar]
  • 18.Zhong Y., Li M., Luan X., Gao F., Wu H., Zi J., Lian Z. Ultrathin ZnIn2S4/ZnSe heteronanosheets with modulated S-scheme enable high efficiency of visible-light-responsive photocatalytic hydrogen evolution. Appl Catal B. 2023;335 doi: 10.1016/j.apcatb.2023.122859. [DOI] [Google Scholar]
  • 19.Liu J., Ma N., Wu W., He Q. Recent progress on photocatalytic heterostructures with full solar spectral responses. Chem. Eng. J. 2020;393 doi: 10.1016/j.cej.2020.124719. [DOI] [Google Scholar]
  • 20.Wang J., Fang Y., Zhang W., Yu X., Wang L., Zhang Y. TiO2/BiOBr 2D–2D heterostructure via in-situ approach for enhanced visible-light photocatalytic N2 fixation. Appl. Surf. Sci. 2021;567 doi: 10.1016/j.apsusc.2021.150623. [DOI] [Google Scholar]
  • 21.Bariki R., Pradhan S.K., Panda S., Nayak S.K., Pati A.R., Mishra B.G. Hierarchical UiO-66(−NH2)/CuInS2 S-Scheme Photocatalyst with Controlled Topology for Enhanced Photocatalytic N2 Fixation and H2O2 Production. Langmuir. 2023;39:7707–7722. doi: 10.1021/acs.langmuir.3c00519. [DOI] [PubMed] [Google Scholar]
  • 22.Esrafili A., Salimi M., Jafari A.J., Sobhi H.R., Gholami M., Kalantary R.R. Pt-based TiO2 photocatalytic systems: A systematic review. J. Mol. Liq. 2022;352 doi: 10.1016/j.molliq.2022.118685. [DOI] [Google Scholar]
  • 23.Lin M., Chen H., Zhang Z., Wang X. Engineering interface structures for heterojunction photocatalysts. PCCP. 2023;25:4388–4407. doi: 10.1039/D2CP05281D. [DOI] [PubMed] [Google Scholar]
  • 24.Guan R., Wang D., Zhang Y., Liu C., Xu W., Wang J., Zhao Z., Feng M., Shang Q., Sun Z. Enhanced photocatalytic N2 fixation via defective and fluoride modified TiO2 surface. Appl. Catal. B. 2021;282 doi: 10.1016/j.apcatb.2020.119580. [DOI] [Google Scholar]
  • 25.Ren G., Shi M., Li Z., Zhang Z., Meng X. Electronic metal-support interaction via defective-induced platinum modified BiOBr for photocatalytic N2 fixation. Appl. Catal. B. 2023;327 doi: 10.1016/j.apcatb.2023.122462. [DOI] [Google Scholar]
  • 26.Li X., Huang Z., Shuck C.E., Liang G., Gogorsi Y., Zhi C. MXene chemistry, electrochemistry and energy storage applications. Nat. Rev. Chem. 2022;6:389–404. doi: 10.1038/s41570-022-00384-8. [DOI] [PubMed] [Google Scholar]
  • 27.Murali G., Modigunta J.K.R., Park Y.H., Lee J.H., Rawal J., Lee S.Y., In I., Park S.J. A Review on MXene Synthesis, Stability, and Photocatalytic Applications. ACS Nano. 2022;16:13370–13429. doi: 10.1021/acsnano.2c04750. [DOI] [PubMed] [Google Scholar]
  • 28.Wang Y., Niu B., Zhang X., Lei Y., Zhong P., Ma X. Review-Ti3C2Tx MXene: An Emerging Two-Dimensional Layered Material in Water Treatment. ECS J. Solid State Sci. Technol. 2021;10(4) doi: 10.1149/2162-8777/abf2de. 047002. [DOI] [Google Scholar]
  • 29.Bury D., Jakubczak M., Purbayanto M.A.K., Wojciechowska A., Moszczyńska D., Jastrzębska A.M. Photocatalytic Activity of the Oxidation Stabilized Ti3C2Tx MXene in Decomposing Methylene Blue, Bromocresol Green and Commercial Textile Dye. Small Methods. 2023:2201252. doi: 10.1002/smtd.202201252. [DOI] [PubMed] [Google Scholar]
  • 30.Li X., Bai Y., Shi X., Su N., Nie G., Zhang R., Nie H., Ye L. Applications of MXene (Ti3C2Tx) in photocatalysis: a review. Mater. Adv. 2021;2:1570–1594. doi: 10.1039/D0MA00938E. [DOI] [Google Scholar]
  • 31.Hussain I., Lamiel C., Javed M.S., Ahmad M., Sahoo S., Chen X., Qin N., Iqbal S., Gu S., Li Y., Chatzichristodoulou C., Zhang K. MXene-based heterostructures: Current trend and development in electrochemical energy storage devices. Prog. Energy Combust. Sci. 2023;97 doi: 10.1016/j.pecs.2023.101097. [DOI] [Google Scholar]
  • 32.Amin I., Brekel H., Nemani K., Batyrev E., Vooys A., Weijde H., Anasori B., Shiju N.R. Ti3C2Tx MXene Polymer Composites for Anticorrosion: An Overview and Perspective. ACS Appl. Mater. Interfaces. 2022;14:43749–43758. doi: 10.1021/acsami.2c11953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Mistry K., Lakhani J.R., Tripathi B., Shinde S., Chandra P. Recent trends in MXene/Metal chalcogenides for electro-/photocatalytic hydrogen evolution reactions. Int. J. Hydrogen Energy. 2022;47(99):41711–41732. [Google Scholar]
  • 34.Huang K., Li C., Meng X. In-situ construction of ternary Ti3C2 MXene@TiO2/ZnIn2S4 composites for highly efficient photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2020;580:669–680. doi: 10.1016/j.jcis.2020.07.044. [DOI] [PubMed] [Google Scholar]
  • 35.Wang K., Li X., Wang N., Shen Q., Liu M., Zhou J., Li N. Z-Scheme Core-Shell meso-TiO2@ZnIn2S4/Ti3C2 MXene Enhances Visible Light-Driven CO2-to-CH4 Selectivity. Ind. Eng. Chem. Res. 2021;60:8720–8732. doi: 10.1021/acs.iecr.1c00713. [DOI] [Google Scholar]
  • 36.Jiang X., Kuklin A.V., Baev A., Ge Y., Agren H., Zhang H., Prasad P.N. Two-dimensional MXenes: From morphological to optical, electric, and magnetic properties and applications. Phys. Rep. 2020;848:1–58. doi: 10.1016/j.physrep.2019.12.006. [DOI] [Google Scholar]
  • 37.Wei S., Fu Y., Liu M., Yue H., Park S., Lee Y.H., Li H., Yao F. Dual-phase MoS2/MXene/CNT ternary nanohybrids for efficient electrocatalytic hydrogen evolution. npj 2D Mater. Appl. 2022;6:25. doi: 10.1038/s41699-022-00300-0. [DOI] [Google Scholar]
  • 38.Du M., Li L., Ji S., Wang T., Wang Y., Zhang J., Li X. In-situ growth of 2D ZnIn2S4 nanosheets on sulfur-doped porous Ti3C2Tx MXene 3D multi-functional architectures for photocatalytic H2 evolution. J. Mater. Chem. C. 2022;10:10636–10644. doi: 10.1039/D2TC01955H. [DOI] [Google Scholar]
  • 39.Ren T., Huang H., Li N., Chen D., Xu Q., Li H., He J., Lu J. 3D hollow MXene@ZnIn2S4 heterojunction with rich zinc vacancies for highly efficient visible-light photocatalytic reduction. J. Colloid Interface Sci. 2021;598:398–408. doi: 10.1016/j.jcis.2021.04.027. [DOI] [PubMed] [Google Scholar]
  • 40.Wang Q., Lei Y., Wang Y., Liu Y., Song C., Zeng J., Song Y., Duan X., Wang D., Li Y. Atomic-scale engineering of chemical-vapor-deposition-grown 2D transition metal dichalcogenides for electrocatalysis. Energ. Environ. Sci. 2020;13:1593–1616. doi: 10.1039/D0EE00450B. [DOI] [Google Scholar]
  • 41.Shi L., Li Z., Ju L., Pena A.C., Orlovskaya N., Zhou H., Yang Y. Promoting nitrogen photofixation over a periodic WS2@TiO2 nanoporous film. J. Mater. Chem. A. 2020;8:1059–1065. doi: 10.1039/C9TA12743G. [DOI] [Google Scholar]
  • 42.Ranjith K.S., Ezhil Vilian A.T., Ghoreishian S.M., Umapathi R., Hwang S.-K., Oh C.W., Huh Y.S., Han Y.-K. Hybridized 1D–2D MnMoO4–MXene nanocomposites as high-performing electrochemical sensing platform for the sensitive detection of dihydroxybenzene isomers in wastewater samples. J. Hazard. Mater. 2022;421:126775. doi: 10.1016/j.jhazmat.2021.126775. [DOI] [PubMed] [Google Scholar]
  • 43.Mositagi S., Ajabshir S.Z., Niasari M.S. Preparation and characterization of BaSnO3 nanostructures vis a new simple surfactant-free route. J. Mater. Sci. Mater. Electron. 2016;27:425–435. doi: 10.1007/s10854-015-3770-0. [DOI] [Google Scholar]
  • 44.Ajabshir S.Z., Morassaei M.S., Amiri O., Niasari M.S., Foong L.K. Nd2Sn2O7 nanostructures: Green synthesis and characterization using date palm extract, a potential electrochemical hydrogen storage material. Ceram. Int. 2020;46:17186–17196. doi: 10.1016/j.ceramint.2020.03.014. [DOI] [Google Scholar]
  • 45.Zhou Y., Wang Y., Wang Y., Li X. Humidity-Enabled Ionic Conductive Trace Carbon Dioxide Sensing of Nitrogen-Doped Ti3C2Tx MXene/Polyethyleneimine Composite Films Decorated with Reduced Graphene Oxide Nanosheets. Anal. Chem. 2020;92(24):16033–16042. doi: 10.1021/acs.analchem.0c03664. [DOI] [PubMed] [Google Scholar]
  • 46.Yang W., Huang B., Li L., Zhang K., Li Y., Huang J., Tang X., Hu T., Yuan K., Chen Y. Covalently Sandwiching MXene by Conjugated Microporous Polymers with Excellent Stability for Supercapacitors. Small Methods. 2020;4:2000434. doi: 10.1002/smtd.202000434. [DOI] [Google Scholar]
  • 47.Miakota D.I., Unocic R.R., Bertoldo F., Ghimire G., Engberg S., Geohegan D., Thygesen K.S., Canulescu S. A facile strategy for the growth of high-quality tungsten disulfide crystals mediated by oxygen-deficient oxide precursors. Nanoscale. 2022;14:9485–9497. doi: 10.1039/D2NR01863B. [DOI] [PubMed] [Google Scholar]
  • 48.Han Y., Liu Y., Su C., Chen X., Li B., Jiang W., Zeng M., Hu N., Su Y., Zhou Z., Zhu Z., Yang Z. Hierarchical WS2–WO3 Nanohybrids with P-N Heterojunctions for NO2 Detection. ACS Appl. Nano Mater. 2021;4:1626–1634. doi: 10.1021/acsanm.0c03094. [DOI] [Google Scholar]
  • 49.Zheng B., Zheng W., Jiang Y., Chen S., Li D., Ma C., Wang X., Huang W., Zhang X., Liu H., Jiang F., Li L., Zhuang X., Wang X., Pan A. WO3-WS2 Vertical Bilayer Heterostructures with High Photoluminescence Quantum Yield. J. Am. Chem. Soc. 2019;141:11754–11758. doi: 10.1021/jacs.9b03453. [DOI] [PubMed] [Google Scholar]
  • 50.Wu Y., Li Y., Li H., Guo H., Yang Q., Li X. Tunning heterostructures interface of Cu2O@HKUST-1 for enhanced photocatalytic degradation of tetracycline hydrochloride. Sep. Purif. Technol. 2022;303 doi: 10.1016/j.seppur.2022.122106. [DOI] [Google Scholar]
  • 51.Keyikoglu R., Doğan I.N., Khataee A., Orooji Y., Kobya M., Yoon Y. Synthesis of visible light responsive ZnCoFe layered double hydroxide towards enhanced photocatalytic activity in water treatment. Chemosphere. 2022;309 doi: 10.1016/j.chemosphere.2022.136534. [DOI] [PubMed] [Google Scholar]
  • 52.Haddadi S., Khataee A., Oskoui S.A., Vahid B., Orooji Y., Yoon Y. Titanium-based MAX-phase with sonocatalytic activity for degradation of oxytetracycline antibiotic. Ultrason. Sonochem. 2023;92 doi: 10.1016/j.ultsonch.2022.106255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Kim S., Nam S.N., Park C.M., Jang M., Qazvini N.T., Yoon Y. Effect of single and multilayered Ti3C2TX MXene as a catalyst and adsorbent on enhanced sonodegradation of diclofenac and verapamil. J. Hazard. Mater. 2022;426 doi: 10.1016/j.jhazmat.2021.128120. [DOI] [PubMed] [Google Scholar]
  • 54.Rad T.S., Yazici E.S., Khataee A., Gengec E., Kobya M. Tuned CuCr layered double hydroxide/carbon-based nanocomposites inducing sonophotocatalytic degradation of dimethyl phthalate. Ultrason. Sonochem. 2023;95 doi: 10.1016/j.ultsonch.2023.106358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Ibragimova R., Erhart P., Rinke P., Komsa H.-P. Surface Functionalization of 2D MXenes: Trends in Distribution, Composition, and Electronic Properties. J. Phys. Chem. Lett. 2021;12:2377–2384. doi: 10.1021/acs.jpclett.0c03710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Yang Y., Ye K., Cao D., Gao P., Qiu M., Liu L., Yang P. Efficient Charge Separation from F– Selective Etching and Doping of Anatase-TiO2{001} for Enhanced Photocatalytic Hydrogen Production. ACS Appl. Mater. Interfaces. 2018;10:19633–19638. doi: 10.1021/acsami.8b02804. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary data 1
mmc1.docx (1.9MB, docx)

Data Availability Statement

Data will be made available on request.


Articles from Ultrasonics Sonochemistry are provided here courtesy of Elsevier

RESOURCES