Next Article in Journal
Synthesis, Structural Properties and Biological Activities of Novel Hydrazones of 2-, 3-, 4-Iodobenzoic Acid
Previous Article in Journal
Study on Calcination Characteristics of Diaspore-Kaolin Bauxite Based on Machine Vision
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of Mesoporous Zr-TiO2 Catalyst with Rich Oxygen Vacancies for N-Methylmorpholine Oxidation to N-Methylmorpholine-N-oxide

1
School of Chemical Engineering and Environment, Weifang University of Science and Technology, Weifang 262700, China
2
Shandong Engineering Research Center of Green and High-Value Marine Fine Chemical, Weifang 262700, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(16), 3812; https://doi.org/10.3390/molecules29163812
Submission received: 17 July 2024 / Revised: 4 August 2024 / Accepted: 9 August 2024 / Published: 11 August 2024

Abstract

:
A series of Zr-TiO2 catalysts were prepared using a facile sol-gel method and were used for N-methylmorpholine (NMM) oxidation to N-methylmorpholine-N-oxide (NMMO). The structure features of Zr-TiO2 catalysts were studied in detail through a variety of characterization methods, such as XRD, SEM, N2 adsorption-desorption isotherms, XPS, EPR, and O2-TPD. As-obtained 5%Zr-TiO2 catalysts had superior catalytic performance and stability with a 97.6% NMMO yield at 40 °C, which related to Zr doping, a higher surface area, more oxygen vacancies, and oxygen chemisorption on the catalytic surface. This work provides an efficient preparation strategy of TiO2-based catalysts for selective oxidation reactions by a facile method.

Graphical Abstract

1. Introduction

Cellulose, an inexhaustible renewable biomass resource in nature, has increasingly caused a great deal of concern in society [1,2,3]. N-methylmorpholine-N-oxide (NMMO) is a highly valuable solvent and a fine chemical for cellulose [4,5,6]. In recent years, most researchers have mainly focused on the application of NMMO in cellulose, with relatively little attention paid to the process of NMMO synthesis. The current method for producing NMMO commercially mainly involves the oxidation conversion reaction of N-methylmorpholine (NMM), in which the catalyst plays an important role [7,8]. Reported catalysts include bicarbonate [9], titanium silicon molecular sieves [10], etc. However, the traditional catalyst has some issues, such as a low catalytic efficiency, high reaction temperature, and many by-products. Therefore, it is still a crucial technological and scientific challenge to develop an efficient and gentle NMM catalytic oxidation system.
In recent years, research activities in the fields of nanoscience and nanotechnology have grown exponentially [11]. In metal oxide nanostructures, TiO2, one of the most popular research materials, is frequently utilized as a catalyst for selective oxidation because of its stability, weak acidity, environmentally friendliness, low cost, and surface properties [12,13,14]. However, the conversion of pure TiO2 may not be sufficiently ideal to meet the needs of commercial applications. The chemical and physical properties of titania can be modified by the doping of metallic ions [15]. The oxygen vacancy can be effectively adjusted by metal doping into TiO2 [16,17]. Nano-scale doping modifications can adjust the TiO2 surface active site, which is expected to further improve the activity of the catalyst while maintaining high selectivity [18]. Therefore, finding more affordable transition metals is a desirable option.
Among numerous doping elements, Zr has received widespread attention due to its unique electronic structure and chemical properties [19,20]. Zr has been used extensively in a variety of acid catalytic processes because of its Lewis acid characteristics under specific reaction circumstances [21,22]. Zr has excellent performance as a catalyst structure additive, can produce strong interactions with the active component, and easily produces oxygen vacancies, which is conducive to improving catalytic activity and selectivity [23,24]. Modifications of TiO2 with oxides of Zr are especially successful in controlling the structure-sorption characteristics of the composites. Zirconium is added during heat treatment at high temperatures to stabilize TiO2, inhibiting the formation of titania crystallites, and increasing the specific surface area [25,26]. Therefore, it is theoretically feasible to design Zr-TiO2 by doping Zr into a TiO2 substrate with good selectivity. Zhao et al. [27] synthesized a series of Zr-TiO2 catalysts, effectively converting ethyl levulinates to gvalerolactone using isopropanol as the H-donor. Kim et al. [28] reported the successful synthesis of Zr-doped TiO2 repeatedly using a solvothermal approach and its application on dye-sensitized solar cells, and achieved good results. These results indicate that metal Zr doping with TiO2 can effectively adjust the crystal structure and surface properties of TiO2, thereby significantly improving its catalytic performance. The study of Zr-TiO2 catalysts has important practical application value. However, there have been no reports on the application of Zr-TiO2 catalysts in NMMO synthesis.
Among the mature chemical synthesis methods, the sol-gel process is one of the common methods for preparing nanomaterials with a high surface area. This method has several advantages, including a simple process, low cost, good dispersion of dopants and titanium sources, and easy adjustments of the catalyst structure and size [29,30]. The emergence of new properties is ensured by mixing components at the molecular level. These are generally predictable and typically rely heavily on factors like the ratio of components, synthesis circumstances, etc. In this work, many materials with varied Zr contents of TiO2 were synthesized using a simple sol-gel technique to enhance the catalytic efficiency of the selective oxidation of NMM with H2O2. A series of characterization techniques were employed to gain a thorough understanding of the catalyst. To attain excellent catalytic performance, many reaction parameters and conditions were methodically tuned. A plausible reaction process for the synthesis of NMMO by NNM oxidation has been proposed, providing a promising method for producing NMMO under mild reaction conditions.

2. Results and Discussion

2.1. Structural Properties of the Catalysts

Figure 1 presents the diffraction patterns of fresh TiO2 and Zr-TiO2. XRD spectra of TiO2 showed the presence of several peaks located at 25.1°, 36.7°, 37.6°, 38.4°, 47.9°, 53.7°, 54.9°, 62.5°, and 68.6°, corresponding to (101), (103), (004), (112), (200), (105), (211), (204), and (116) crystallographic planes of anatase, respectively, which are in agreement with the anatase TiO2 standard card (JCPDS No. 21-1272) [31]. Following zirconium doping, the XRD patterns of TiO2 resembled those of pure TiO2, and no peaks corresponding to zirconium oxides were seen. This suggests that the zirconium element may be present on the surface of the catalyst in the form of highly dispersed zirconia [32,33]. Zr4+ is more electropositive than Ti4+; therefore, each TiO2 nanoparticle may have a loosely held electron cloud, which promotes the creation of a less dense anatase phase [34]. Furthermore, the XRD patterns show that the positions of the XRD peaks did not change noticeably as the amount of zirconium doping increased. However, the fact that the Zr-TiO2 XRD peaks show considerable peak broadening in comparison to pure TiO2 suggests that the particle size of TiO2 doped with Zr element is smaller.
Table 1 displays the average particle size of the catalyst, which was calculated based on the (101) peak of the anatase using the Scherrer equation:
D = K λ β c o s θ
where D is the average crystallite size of the catalyst, λ is the X-ray wave-length (1.54 Å), β is the full width at half maximum (FWHM) of the sample, K = 0.89, and θ is the diffraction angle. The average particle size of the catalyst decreases first and then increases weakly with increasing zirconium doping on TiO2 (Table 1). Therefore, the proper doping amount can prevent the growth and aggregation of TiO2 particles.
Next, SEM, TEM, and EDS studies were used to examine the morphology and elemental composition of the 5%Zr-TiO2 (Figure 2). SEM images (Figure 2a) revealed that the sample is composed of uniform small particles with no fixed morphology [35]. TEM images (Figure 2b,c) further confirmed that the 5%Zr-TiO2 powder was composed of nanocrystalline grains, which are evenly distributed with an average particle size of 10–20 nm. This result is consistent with the particle size calculated from XRD analysis data. Additionally, Figure 2d displays the particle distribution of different crystal surfaces. Two TiO2 crystal surfaces (101) and (004) with lattice values of 0.351 and 0.238 nm are observed [36]. Further, there is a crystal surface of d = 0.286 nm, which was assigned to ZrO2 (111), indicating that the doped Zr element in TiO2 exists mainly in the state of Zr oxide [37,38]. Moreover, EDS mapping (Figure 2e–h) showed the presence of O, Ti, and Zr as primary elements. This indicated that the doped Zr elements were uniformly distributed on the surface of the catalyst. These tests validated the successful integration of Zr species, and the efficacy of the synthetic methodology that was designed.
As shown in Figure 3 and Table 1, the specific surface area and aperture information of catalysis were determined. Figure 3a shows that these catalysts have a hysteresis loop in the N2 adsorption-desorption isotherm. The isotherm displayed in all samples is a classic IV type, representing porous solids. The hysteresis loop indicates the presence of ink bottle or ink cage holes [15,39]. Moreover, most of the aperture distribution is concentrated in the 2–10 nm region (Figure 3b). Notably, Zr-TiO2 has a smaller pore size and larger specific surface area than pure TiO2 (Table 1). Additionally, a higher surface area was observed in the 5%Zr-TiO2 catalyst with a smaller crystal size. This can expose more active sites and improve the catalytic performance of NMM oxidation. The findings suggest that adding Zr to TiO2 species can modify the structural network to provide improved porous surfaces. This would increase the number of sites available for catalytic reactions and be advantageous for reactant molecule adsorption [40].
XPS examination provided additional information about the chemical state and surface compositions of the catalysts (Figure 4 and Table 2). With reference to the C 1s peak position, which is fixed at 284.8 eV, the shift of the conventional peaks represented by solid curves was calibrated. The Zr element signal in the full spectrum of XPS gradually increased with the increase of Zr content (Figure S1). Two prominent peaks were observed in the XPS spectrum in the Ti 2p area for all catalysts in Figure 4a. These peaks are associated with the Ti 2p3/2 and Ti 2p1/2 signals of Ti species, and are situated at 458.4 and 464.0 eV, respectively, implying that Ti basically forms as Ti4+ [26,41]. Furthermore, two peaks were identified in the Zr 3d XPS region of the catalysts, which were linked to the Zr 3d5/2 and Zr 3d3/2 signals, respectively, and are positioned at 181.6 and 184.0 eV (Figure 4b). Furthermore, the separation of Zr 3d5/2 and Zr 3d3/2 signals was measured and found to be roughly 2.4 eV, which means that Zr exists chiefly in the Zr4+ state [42,43]. As the Zr content increased, there was a slight shift in the binding energy of Zr 3d towards a higher binding energy, which may be attributed to the interaction between Zr and Ti [44].
Furthermore, two distinct contributions were seen in the O1s XPS area (Figure 4c) at about 529.6 and 531.2 eV. These contributions are associated with the lattice oxygen (Olatt) in bulk TiO2 and surface adsorbed oxygen (Oads), which was ascribed to Ti-O and surface -OH bonding [45,46]. With the increase of Zr content, the Oads ratio of Zr-TiO2 first increased and then decreased. In contrast, Table 2 indicates that the 5%Zr-TiO2 catalyst has the highest proportion of Oads. This is likely due to the presence of more oxygen-defective sites, which have the tendency to interact with hydrogen atoms and generate more hydroxyl surface groups [47,48]. Moreover, Table 2 provides an overview of comprehensive information regarding element analysis. The Zr/Ti ratios of the surface are less than the Zr/Ti ratios predicted by theory, indicating an enrichment of Ti4+ on the catalyst surface. The O/Ti ratios of the surface are lower than the theoretical O/Ti ratios, indicating that the catalyst has surface defects and oxygen vacancies, which is helpful for improving catalytic qualities [49]. The C/Ti ratio of the sample surface is about 0.7–0.8, which may be attributed to the contact of the sample surface with carbon dioxide or other organic substances in the air, resulting in a carbon-containing signal in the test results.
O2-TPD was used to characterize the oxygen species of the catalysts, and the findings are displayed in Figure 5a. Within the temperature range of 0–800 °C, three different types of desorption peaks were detected in the sample. These peaks occur at approximately 220, 425, and 600 °C. Broadly speaking, there were three types of desorbed oxygen species found in the catalysts: adsorbed oxygen (O2 and O), lattice oxygen (O2−), and bulk lattice oxygen (Obulk2−) [50,51]. The surface of each catalyst is weakly linked to the chemically adsorbed oxygen species O2 and O, for which desorption took place between 150 and 350 °C. The desorption temperature range for the surface O2− was between 350 and 670 °C, which is higher than the bulk lattice oxygen’s desorption temperature of 700 °C [52]. Since surface lattice oxygen desorbs at a temperature over 350 °C, the catalyst’s peak below 350 °C represents the desorption of chemically adsorbed oxygen [53,54]. NMM typically oxidizes at temperatures lower than 100 °C, with surface adsorbed oxygen taking part in the catalytic oxidation reaction. Figure 5a shows that the Zr-TiO2 catalysts have a much larger O2 desorption capacity than the pure TiO2 catalyst, particularly the 5%Zr-TiO2 catalyst. This suggests that adding Zr to the catalyst increases the number of active oxygen species. To further study the oxygen vacancy signal on the catalyst surface, oxygen vacancy measurements were recorded by paramagnetic resonance spectrometer (Figure 5b). As the Zr content increased, the oxygen vacancy signal was obviously enhanced. By maximizing the reactant adsorption energy on the catalyst surface, oxygen vacancies can lower the reaction energy barrier and encourage molecular activation. In the catalyst, oxygen vacancies collaborate with adjacent active metal sites in a beneficial manner [55]. As a result, the addition of Zr species accelerated the oxidation of NMM by creating more oxygen vacancies and surface oxygen.

2.2. Catalytic Performance

In our preliminary research, the metal oxides CuO, MgO, ZnO, Al2O3, ZrO2, TiO2, and SiO2 were proven to be active for NMM oxidation (Figure S2). As expected, we found that TiO2 exhibits high catalytic performance in the oxidation of NMM to NMMO, with excellent NMMO selectivity. Furthermore, ZrO2 shows slightly better conversion and slightly lower selectivity than TiO2.
In this study, to compare the performance of the prepared catalyst, catalyst activity screening experiments were carried out by magnetic stirring in a three-neck flask at 30 °C for 0.5 h. The prepared catalyst was tested during the oxidation process of NMM by H2O2, and the results are shown in Figure 6a. The statistical results of these repeated experiments are summarized in Table S1.
As can be seen from Figure 6a, pure TiO2 provided only 35.5% NMM conversion and approximately 99.8% NMMO selectivity. As expected, when the Zr content in TiO2 increased from 1% to 10%, the NMM conversion first increased and then decreased, with 5%Zr-TiO2 showing the highest conversion rate (51.4%). Under the same circumstances, the NMMO selectivity showed a weak trend of continuously decreasing. This suggested that a key factor in the catalytic oxidation process was the interaction between the appropriate amount of Zr and Ti structures. Clearly, 5%Zr-TiO2 showed a slightly higher NMMO yield. Consequently, 5%Zr-TiO2 was determined to be the ideal catalyst for additional tuning based on NMMO yields.
Figure 6b–e illustrates the results of a study conducted on the impacts of reaction parameters, including temperature, reaction duration, catalyst dosage, and H2O2 dosage. Increasing the temperature led to higher conversions, but a temperature above 40 °C caused a noticeable decrease in NMMO yield due to reduced selectivity (Figure 6b). A possible explanation for this result is that high temperature is conductive to the decomposing of H2O2 into free radicals, which results in the formation of by-products [56]. Therefore, the active oxygen in H2O2 is not utilized effectively under high temperatures. Increasing the reaction time gradually increased the conversion, but when the reaction time exceeded 3 h, the selectivity decreased significantly (Figure 6c). The NMM conversion rose as the catalyst dosage increased from 10 mg to 40 mg, although the NMMO yield first climbed and subsequently somewhat dropped due to a decrease in selectivity (Figure 6d). Therefore, the optimal catalyst dosage is 20 mg. In a similar manner, the oxidizability of NMM was investigated at various molar ratios of H2O2/NMM (Figure 6e). A high molar ratio of H2O2 to substrate is advantageous for selective oxidation within a certain range. Increased H2O2 dosages result in increased conversions; however, if the dose is greater than 0.13 mol, the over-oxidation of NMM causes a noticeable drop in yield. Therefore, the optimum H2O2 dosage is 0.13 mol. Based on the previously provided information, the optimal reaction conditions were determined to be 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 40 °C, and 3 h. Under optimum reaction circumstances, the conversion of NMM and the selectivity of NMMO can approach 98.7% and 98.9%, respectively, even with a minimal amount of catalyst.
In terms of cost-effective and eco-friendly procedures, the durability and recyclability of catalysts are essential elements. Following the reaction, the catalyst was filtered; cleaned three times with ethanol, water, and dichloromethane; and then dried for 3 h at 80 °C. After the sixth run, the NMM conversion barely decreased to 97.6%, with a slight decrease in selectivity (Figure 6f). The findings showed that, under ideal circumstances, the 5%Zr-TiO2 combination was very stable in this investigation. A comparison between the performance of the catalyst prepared in this study and the traditional catalyst is summarized in Table S2.

2.3. Reaction Process

The XRD and TEM data demonstrated that the sol-gel approach is an efficient means to create Zr-TiO2 catalysts with a variety of crystal sizes. The Zr content of these catalysts can be adjusted to alter their crystal sizes. Smaller crystal diameters are typically the consequence of increased Zr concentrations. For catalysts with smaller crystal sizes, the N2 adsorption-desorption isotherms show a greater surface area. Zr doping increases the amount of adsorbed oxygen and improves surface active sites, as confirmed by XPS analysis, which also shows the presence of surface defects and oxygen vacancies. The results of O2-TPR and EPR tests indicate that the 5%Zr-TiO2 catalyst with more adsorbed oxygen and oxygen vacancies has superior catalytic oxidation, producing more active oxygen species and increasing the catalytic oxidation of NMM.
Based on the above data and free radical test analysis (Figure S3), a plausible reaction process for the selective catalytic oxidation of NMM to NMMO by Zr-TiO2 with H2O2 was delineated in Scheme 1. The catalytic activity of Zr-TiO2 is enhanced by its large specific surface area structure and plentiful surface oxygen vacancies. First, some H2O2 is adsorbed onto the catalyst surface for activation, forming a metal peroxide species [57] that improves H2O2 utilization efficiency by selectively oxidizing NMM to NMMO. In the meantime, activation of H2O2 produces strong oxidizing hydroxyl radicals (•OH) and peroxide radicals (HO2•) [58], which function as active oxygen species to oxidize NMM to NMMO with minimal by-products. Additionally, reactive oxygen species (O2 and O) are formed by the catalyst’s surface chemical adsorption activation [51], which oxidizes NMM to NMMO and speeds up the oxidation of the substrate NMM.

3. Experimental Section

3.1. Chemical Reagents

Butyl titanate, zirconium nitrate pentahydrate, anhydrous ethanol, glacial acetic acid, and dichloromethane were purchased from Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). Every chemical used was of p.a. quality, and no further purification was necessary.

3.2. Zr-TiO2 Catalyst Preparation

A series of Zr-TiO2 nanoparticles were prepared by the sol-gel method. Solution A was prepared by adding 17 g (50 mmol) of butyl titanate dropwise to 20 mL of anhydrous ethanol and stirring for 30 min. Solution B was prepared by dissolving zirconium nitrate pentahydrate (x mmol) in a homogeneous solution of anhydrous ethanol (10 mL), glacial acetic acid (4 mL), and deionized water (2 mL) and stirring for 30 min. Then, solution B was added dropwise to solution A. The pH of the solution was adjusted to 2 using acetic acid and stirred for 6 h to form a sol. Subsequently, sonication was performed for 30 min, followed by aging at 50 °C for 12 h. Zr-TiO2 nanopowder was obtained through vacuum drying and calcination at 500 °C for 3 h. The samples with different amounts of zirconium nitrate pentahydrate added (x = 0, 0.5, 1.5, 2.5, 3.5, 5.0 mmol) were denoted as TiO2, 1%Zr-TiO2, 3%Zr-TiO2, 5%Zr-TiO2, 7%Zr-TiO2, and 10%Zr-TiO2.

3.3. Characterization

Powder XRD (D8 Advance A25, Bruker, Bremen, Germany) was used to analyze the crystal structure and arrangement of the sample throughout a 2θ range of 10° to 80°. The morphology and the microstructure of the Zr-TiO2 nanopowder were examined by SEM (MIRA LMS, TESCAN, Brno, Czech Republic) and TEM (Talos F200X G2, FEI, Waltham, MA, USA). The surface chemical composition and elemental state were analyzed by X-ray photoelectron spectroscopy (Scientific K-Alpha, Thermo, Waltham, MA, USA). The specific surface area information was measured by a N2 adsorption and desorption isotherm instrument (ASAP 2460, Micromeritics, Atlanta, GA, USA). The specific surface area was determined by the Brunauer–Emmett–Teller (BET) method and the pore size was obtained by the Barrett–Joyner–Halenda (BJH) method. O2 temperature-programmed desorption was performed under a chemisorption analyzer (5080B, xianquan, Tianjin, China). Free radical and oxygen vacancy measurements were recorded by paramagnetic resonance spectrometer (EMXnano, Bruker, Bremen, Germany).

3.4. Catalytic Activity Test

Quantitative NMM (0.1 mol) and a trace amount of catalyst (20 mg) were added to a three-necked jacketed round bottom flask for the NMM oxidation reaction. At 30 °C, 0.13 mol H2O2 was gradually added to the NMM/catalyst mixture while being stirred by magnetic means. Following the injection of H2O2, the reaction system was maintained at 30 °C. Reaction product samples were taken every 0.5 h for HPLC (Flexar, PerkinElmer, MC, USA) analysis to ascertain the contents of NMM and NMMO. During the test, morpholine (M) and nitroso-morpholine (NMOR) were mainly considered as possible by-products in the reaction system. Based on this, standard curves were established using standard samples of M, NMM, NMMO, and NMOR (Figure S4).
The conversion and selectivity were calculated according to Equations (1) and (2):
C o n v e r s i o n   ( % ) = C ( N M M ) I n i t i a l C ( N M M ) F i n a l C ( N M M ) I n i t i a l × 100
S e l e c t i v i t y   ( % ) = C ( N M M O ) C ( N M M ) I n i t i a l C ( N M M ) F i n a l × 100
where C(NMM) and C(NMMO) represent the concentrations of NMM and product NMMO (mol/mL). Throughout the studies, it was expected that the system’s overall volume would not change. The conversion and selectivity figures were all given as molar percentages.

4. Conclusions

In a word, a series of efficient Zr-TiO2 catalysts were prepared by a facile sol-gel method and applied to NMM oxidation to NMMO. An appropriate quantity of the transition metal Zr can be added to the catalyst to effectively increase its specific surface area and the contact area of multiphase catalytic processes, according to experiments and characterizations. Simultaneously, the addition of Zr considerably raises the oxygen vacancy concentration and oxygen chemisorption on the catalytic surface, which greatly increases the activity of the catalyst. The catalytic activity experiment indicated that 5%Zr-TiO2 exhibited the best catalytic performance, and the optimized conditions for the reaction were determined to be 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 40 °C, and 3 h. Even with a small amount of catalyst, the conversion of NMM and yield of NMMO reached 98.7% and 97.6%, respectively. In addition, cyclic experiments showed that 5%Zr-TiO2 catalysts have good stability. In conclusion, this study offered a new strategy for the design of highly efficient catalysts in NMMO industrial production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29163812/s1. Figure S1: XPS full spectra of the catalysts. Figure S2: Performance comparison of different catalysts. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg catalysts, 30 °C, 0.5 h. Table S1: Effect of different catalysts on NMM oxidation by H2O2. Figure S3: EPR spectra of 5%Zr-TiO2. Figure S4: HPLC standard curve. Table S2: Comparison with reported catalyst properties.

Author Contributions

Conceptualization, Y.L.; Investigation, Y.L. and Z.F.; Data curation, Y.L., L.F., J.L. and C.Z.; Writing—original draft, Y.L.; Writing—review and editing, Y.L., L.F., F.L. and J.L.; Visualization, Y.S.; Supervision, Z.F. and F.L. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by Natural Science Foundation of Shandong Provincial (No. ZR2022MB118).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding authors.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (No. 52301100), Natural Science Foundation of Shandong Province (No. ZR2022MB118), and Weifang University of Science and Technology Program (No. KJRC2023022).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Aziz, T.; Farid, A.; Haq, F.; Kiran, M.; Ullah, A.; Zhang, K.; Li, C.; Ghazanfar, S.; Sun, H.; Ullah, R. A review on the modification of cellulose and its applications. Polymers 2022, 14, 3206. [Google Scholar] [CrossRef] [PubMed]
  2. Etale, A.; Onyianta, A.J.; Turner, S.R.; Eichhorn, S.J. Cellulose: A review of water interactions, applications in composites, and water treatment. Chem. Rev. 2023, 123, 2016–2048. [Google Scholar] [CrossRef] [PubMed]
  3. D’Acierno, F.; Michal, C.A.; MacLachlan, M.J. Thermal stability of cellulose nanomaterials. Chem. Rev. 2023, 123, 7295–7325. [Google Scholar] [CrossRef]
  4. Agarwala, P.; Kanthale, P.; Thakre, S. Flocculation of spin bath NMMO solution for removal of colloidal impurities in lyocell process. Sep. Sci. Technol. 2019, 54, 2516–2527. [Google Scholar] [CrossRef]
  5. Jadhav, S.; Lidhure, A.; Thakre, S.; Ganvir, V. Modified Lyocell process to improve dissolution of cellulosic pulp and pulp blends in NMMO solvent. Cellulose 2021, 28, 973–990. [Google Scholar] [CrossRef]
  6. Oliva, A.; Tan, L.; Papirio, S.; Esposito, G.; Lens, P. Use of N-Methylmorpholine-N-oxide (NMMO) pretreatment to enhance the bioconversion of ligno-cellulosic residues to methane. Biomass Convers. Biorefin. 2024, 14, 11113–11130. [Google Scholar] [CrossRef] [PubMed]
  7. Markosyan, A.; Baghdasaryan, G.; Hovhannisyan, G.; Attaryan, H.; Hasratyan, G. Waste-free technology for N-methylmorpholine synthesis. Russ. J. Appl. Chem. 2013, 86, 845–847. [Google Scholar] [CrossRef]
  8. El-Wakil, N.; Taha, M.; Abouzeid, R.; Dufresne, A. Dissolution and regeneration of cellulose from N-methylmorpholine N-oxide and fabrication of nanofibrillated cellulose. Biomass Convers. Biorefin. 2024, 14, 5399–5410. [Google Scholar] [CrossRef]
  9. Balagam, B.; Richardson, D.E. The mechanism of carbon dioxide catalysis in the hydrogen peroxide N-oxidation of amines. Inorg. Chem. 2008, 47, 1173–1178. [Google Scholar] [CrossRef]
  10. Prasad, M.R.; Kamalakar, G.; Madhavi, G.; Kulkarni, S.; Raghavan, K. An efficient synthesis of heterocyclic N-oxides over molecular sieve catalysts. J. Mol. Catal. A 2002, 186, 109–120. [Google Scholar] [CrossRef]
  11. Bouzakher-Ghomrasni, N.; Tache, O.; Leroy, J.; Feltin, N.; Testard, F.; Chivas-Joly, C. Dimensional measurement of TiO2 (Nano) particles by SAXS and SEM in powder form. Talanta 2021, 234, 122619. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, L.; Chen, X. Titanium dioxide nanomaterials: Self-structural modifications. Chem. Rev. 2014, 114, 9890–9918. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, Y.; Chen, S.; Chang, X.; Wang, X.; Sun, G.; Lu, Z.; Zhao, Z.-J.; Pei, C.; Gong, J. Ultrathin TiOx nanosheets rich in tetracoordinated Ti sites for propane dehydrogenation. ACS Catal. 2023, 13, 6104–6113. [Google Scholar] [CrossRef]
  14. Li, W.; Wu, Z.; Wang, J.; Elzatahry, A.A.; Zhao, D. A perspective on mesoporous TiO2 materials. Chem. Mater. 2014, 26, 287–298. [Google Scholar] [CrossRef]
  15. Bineesh, K.V.; Kim, D.-K.; Park, D.-W. Synthesis and characterization of zirconium-doped mesoporous nano-crystalline TiO2. Nanoscale 2010, 2, 1222–1228. [Google Scholar] [CrossRef] [PubMed]
  16. Xiao, L.; Xie, Z.; Song, S.; Zhao, Z.; Ke, M.; Song, W.; Zhao, Z.; Liu, J. Descriptor-guided design and experimental synthesis of metal-doped TiO2 for propane dehydrogenation. Ind. Eng. Chem. Res. 2021, 60, 1200–1209. [Google Scholar] [CrossRef]
  17. Li, J.; Jin, Z.; Zhang, Y.; Liu, D.; Ma, A.; Sun, Y.; Li, X.; Cai, Q.; Gui, J. Ag-induced anatase-rutile TiO2−x heterojunction facilitating the photogenerated carrier separation in visible-light irradiation. J. Alloys Compd. 2022, 909, 164815. [Google Scholar] [CrossRef]
  18. Wang, C.-T.; Lin, J.-C. Surface nature of nanoparticle zinc-titanium oxide aerogel catalysts. Appl. Surf. Sci. 2008, 254, 4500–4507. [Google Scholar] [CrossRef]
  19. Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-based metal-organic frameworks: Design, synthesis, structure, and applications. Chem. Soc. Rev. 2016, 45, 2327–2367. [Google Scholar] [CrossRef]
  20. Song, J.; Wu, L.; Zhou, B.; Zhou, H.; Fan, H.; Yang, Y.; Meng, Q.; Han, B. A new porous Zr-containing catalyst with a phenate group: An efficient catalyst for the catalytic transfer hydrogenation of ethyl levulinate to γ-valerolactone. Green Chem. 2015, 17, 1626–1632. [Google Scholar] [CrossRef]
  21. Yang, J.; Huang, W.; Liu, Y.; Zhou, T. Enhancing the conversion of ethyl levulinate to γ-valerolactone over Ru/UiO-66 by introducing sulfonic groups into the framework. RSC Adv. 2018, 8, 16611–16618. [Google Scholar] [CrossRef] [PubMed]
  22. Akram, M.; Javed, H.M.A.; Ghaffar, A.; Majeed, M.I. Investigations of Zr/TiO2-based nanocomposites for efficient plasmonic dye-sensitized solar cells. Int. J. Mod. Phys. B 2023, 37, 2350042. [Google Scholar] [CrossRef]
  23. Tabanelli, T.; Paone, E.; Blair Vásquez, P.; Pietropaolo, R.; Cavani, F.; Mauriello, F. Transfer hydrogenation of methyl and ethyl levulinate promoted by a ZrO2 catalyst: Comparison of batch vs continuous gas-flow conditions. ACS Sustain. Chem. Eng. 2019, 7, 9937–9947. [Google Scholar] [CrossRef]
  24. Xue, Z.; Jiang, J.; Li, G.; Zhao, W.; Wang, J.; Mu, T. Zirconium-cyanuric acid coordination polymer: Highly efficient catalyst for conversion of levulinic acid to γ-valerolactone. Catal. Sci. Technol. 2016, 6, 5374–5379. [Google Scholar] [CrossRef]
  25. Kim, J.; Song, K.C.; Foncillas, S.; Pratsinis, S.E. Dopants for synthesis of stable bimodally porous titania. J. Eur. Ceram. Soc. 2001, 21, 2863–2872. [Google Scholar] [CrossRef]
  26. Gnatyuk, Y.; Smirnova, N.; Korduban, O.; Eremenko, A. Effect of zirconium incorporation on the stabilization of TiO2 mesoporous structure. Surf. Interface Anal. 2010, 42, 1276–1280. [Google Scholar] [CrossRef]
  27. Zhao, D.; Su, T.; Rodríguez-Padrón, D.; Lü, H.; Len, C.; Luque, R.; Yang, Z. Efficient transfer hydrogenation of alkyl levulinates to γ-valerolactone catalyzed by simple Zr-TiO2 metal oxide systems. Mater. Today Chem. 2022, 24, 100745. [Google Scholar] [CrossRef]
  28. Kim, S.-J.; Kang, M.-S. Synthesis of Zr-incorporated TiO2 using a solvothermal method and its photovoltaic efficiency on dye-sensitized solar cells. Bull. Korean Chem. Soc. 2011, 32, 3317–3322. [Google Scholar] [CrossRef]
  29. Bokov, D.; Turki Jalil, A.; Chupradit, S.; Suksatan, W.; Javed Ansari, M.; Shewael, I.H.; Valiev, G.H.; Kianfar, E. Nanomaterial by sol-gel method: Synthesis and application. Adv. Mater. Sci. Eng. 2021, 2021, 5102014. [Google Scholar] [CrossRef]
  30. Parashar, M.; Shukla, V.K.; Singh, R. Metal oxides nanoparticles via sol-gel method: A review on synthesis, characterization and applications. J. Mater. Sci. Mater. Electron. 2020, 31, 3729–3749. [Google Scholar] [CrossRef]
  31. Chen, X.; Liu, L.; Yu, P.Y.; Mao, S.S. Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 2011, 331, 746–750. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, H.; Liu, G.; Zhou, Q. Preparation and characterization of Zr doped TiO2 nanotube arrays on the titanium sheet and their enhanced photocatalytic activity. J. Solid State Chem. 2009, 182, 3238–3242. [Google Scholar] [CrossRef]
  33. Chen, X.; Wang, X.; Fu, X. Hierarchical macro/mesoporous TiO2/SiO2 and TiO2/ZrO2 nano-composites for environmental photocatalysis. Energy Environ. Sci. 2009, 2, 872–877. [Google Scholar] [CrossRef]
  34. Venkatachalam, N.; Palanichamy, M.; Arabindoo, B.; Murugesan, V. Enhanced photocatalytic degradation of 4-chlorophenol by Zr4+ doped nano TiO2. J. Mol. Catal. A Chem. 2007, 1, 158–165. [Google Scholar] [CrossRef]
  35. Gerasimova, T.; Evdokimova, O.; Kraev, A.; Ivanov, V.; Agafonov, A. Micro-mesoporous anatase TiO2 nanorods with high specific surface area possessing enhanced adsorption ability and photocatalytic activity. Microporous Mesoporous Mater. 2016, 235, 185–194. [Google Scholar] [CrossRef]
  36. Thompson, T.L.; Yates, J.T. Surface science studies of the photoactivation of TiO2 new photo-chemical processes. Chem. Rev. 2006, 106, 4428–4453. [Google Scholar] [CrossRef]
  37. Qian, J.; Hou, X.; Qin, Z.; Li, B.; Tong, Z.; Dong, L.; Dong, L. Enhanced catalytic properties of Cu-based composites for NOx reduction with coexistence and intergrowth effect. Fuel 2018, 234, 296–304. [Google Scholar] [CrossRef]
  38. Qian, J.; Hu, Q.; Hou, X.; Qian, F.; Dong, L.; Li, B. Study of different Ti/Zr ratios on the physicochemical properties and catalytic activities for CuO/Ti-Zr-O composites. Ind. Eng. Chem. Res. 2018, 57, 12792–12800. [Google Scholar] [CrossRef]
  39. Suib, S.L. A review of recent developments of mesoporous materials. Chem. Rec. 2017, 17, 1169–1183. [Google Scholar] [CrossRef]
  40. Suttiponparnit, K.; Jiang, J.; Sahu, M.; Suvachittanont, S.; Charinpanitkul, T.; Biswas, P. Role of surface area, primary particle size, and crystal phase on titanium dioxide nanoparticle dispersion properties. Nanoscale Res. Lett. 2011, 6, 27. [Google Scholar] [CrossRef]
  41. Huang, J.; Kang, Y.; Wang, L.; Yang, T.; Wang, Y.; Wang, S. Mesoporous CuO/TixZr1−xO2 catalysts for low-temperature CO oxidation. Catal. Commun. 2011, 15, 41–45. [Google Scholar] [CrossRef]
  42. Tsunekawa, S.; Asami, K.; Ito, S.; Yashima, M.; Sugimoto, T. XPS study of the phase transition in pure zirconium oxide nanocrystallites. Appl. Surf. Sci. 2005, 252, 1651–1656. [Google Scholar] [CrossRef]
  43. Nelson, A.E.; Schulz, K.H. Surface chemistry and microstructural analysis of CexZr1−xO2−y model catalyst surfaces. Appl. Surf. Sci. 2003, 210, 206–221. [Google Scholar] [CrossRef]
  44. Duan, B.; Zhou, Y.; Huang, C.; Huang, Q.; Chen, Y.; Xu, H.; Shen, S. Impact of Zr-doped TiO2 photocatalyst on formaldehyde degradation by Na addition. Ind. Eng. Chem. Res. 2018, 57, 14044–14051. [Google Scholar] [CrossRef]
  45. Zhang, J.; Li, L.; Liu, D.; Zhang, J.; Hao, Y.; Zhang, W. Multi-layer and open three-dimensionally ordered macroporous TiO2-ZrO2 composite: Diversified design and the comparison of multiple mode photocatalytic performance. Mater. Des. 2015, 86, 818–828. [Google Scholar] [CrossRef]
  46. Divya, G.; Jaishree, G.; Sivarao, T.; Lakshmi, K.D. Microwave assisted sol-gel approach for Zr doped TiO2 as a benign photocatalyst for bismark brown red dye pollutant. RSC Adv. 2023, 13, 8692–8705. [Google Scholar] [CrossRef] [PubMed]
  47. Gurubasavaraj, P.M.; Roesky, H.W.; Sharma, P.M.V.; Oswald, R.B.; Dolle, V.; Herbst-Irmer, R.; Pal, A. Oxygen Effect in Heterobimetallic Catalysis: The Zr-O-Ti System as an Excellent Example for Olefin Polymerization. Organometallics 2007, 26, 3346–3351. [Google Scholar] [CrossRef]
  48. Chen, J.; Ding, Z.; Wang, C.; Hou, H.; Zhang, Y.; Wang, C.; Zou, G.; Ji, X. Black anatase titania with ultrafast sodium-storage performances stimulated by oxygen vacancies. ACS Appl. Mater. Interfaces 2016, 8, 9142–9151. [Google Scholar] [CrossRef]
  49. Ma, J.; Jin, G.; Gao, J.; Li, Y.; Dong, L.; Huang, M.; Huang, Q.; Li, B. Catalytic effect of two-phase intergrowth and coexistence CuO-CeO2. J. Mater. Chem. A 2015, 3, 24358–24370. [Google Scholar] [CrossRef]
  50. Liu, Y.; Luo, M.; Wei, Z.; Xin, Q.; Ying, P.; Li, C. Catalytic oxidation of chlorobenzene on supported manganese oxide catalysts. Appl. Catal. B Environ. 2001, 29, 61–67. [Google Scholar] [CrossRef]
  51. Hu, D.; Li, W.; Yin, K.; Huang, B. Promoting effect of Ru-doped Mn/TiO2 catalysts for catalytic oxidation of chlorobenzene. New J. Chem. 2022, 46, 10820–10828. [Google Scholar] [CrossRef]
  52. Song, W.; Poyraz, A.S.; Meng, Y.; Ren, Z.; Chen, S.-Y.; Suib, S.L. Mesoporous Co3O4 with controlled porosity: Inverse micelle synthesis and high-performance catalytic CO oxidation at −60 °C. Chem. Mater. 2014, 26, 4629–4639. [Google Scholar] [CrossRef]
  53. Chen, X.; Cai, S.; Yu, E.; Li, J.; Chen, J.; Jia, H. Photothermocatalytic performance of ACo2O4 type spinel with light-enhanced mobilizable active oxygen species for toluene oxidation. Appl. Surf. Sci. 2019, 484, 479–488. [Google Scholar] [CrossRef]
  54. Feng, X.; Tian, M.; He, C.; Li, L.; Shi, J.-W.; Yu, Y.; Cheng, J. Yolk-shell-like mesoporous CoCrOx with superior activity and chlorine resistance in dichloromethane destruction. Appl. Catal. B Environ. 2020, 264, 118493. [Google Scholar] [CrossRef]
  55. Beniya, A.; Higashi, S. Towards dense single-atom catalysts for future automotive applications. Nat. Catal. 2019, 2, 590–602. [Google Scholar] [CrossRef]
  56. Yang, X.J.; Tian, P.F.; Wang, H.L.; Xu, J.; Han, Y.F. Catalytic decomposition of H2O2 over a Au/carbon catalyst: A dual intermediate model for the generation of hydroxyl radicals. J. Catal. 2016, 336, 126–132. [Google Scholar] [CrossRef]
  57. Zhong, Y.; Liang, X.; He, Z.; Tan, W.; Zhu, J.; Yuan, P.; Zhu, R.; He, H. The constraints of transition metal substitutions (Ti, Cr, Mn, Co and Ni) in magnetite on its catalytic activity in heterogeneous Fenton and UV/Fenton reaction: From the perspective of hydroxyl radical generation. Appl. Catal. B Environ. 2014, 150, 612–618. [Google Scholar] [CrossRef]
  58. Sun, G.; Li, M.M.-J.; Nakagawa, K.; Li, G.; Wu, T.-S.; Peng, Y.-K. Bulk-to-nano regulation of layered metal oxide gears H2O2 activation pathway for its stoichiometric utilization in selective oxidation reaction. Appl. Catal. B Environ. 2022, 313, 121461. [Google Scholar] [CrossRef]
Figure 1. XRD pattern comparison of the catalysts.
Figure 1. XRD pattern comparison of the catalysts.
Molecules 29 03812 g001
Figure 2. SEM (a), TEM (bd), and EDS mapping of Zr, Ti, and O element (eh) of 5%Zr-TiO2.
Figure 2. SEM (a), TEM (bd), and EDS mapping of Zr, Ti, and O element (eh) of 5%Zr-TiO2.
Molecules 29 03812 g002
Figure 3. (a) N2 adsorption-desorption isothermal profiles and (b) pore size distribution of the catalysts.
Figure 3. (a) N2 adsorption-desorption isothermal profiles and (b) pore size distribution of the catalysts.
Molecules 29 03812 g003
Figure 4. XPS spectra of all catalysts: (a) Ti 2p, (b) Zr 3d, and (c) O 1s.
Figure 4. XPS spectra of all catalysts: (a) Ti 2p, (b) Zr 3d, and (c) O 1s.
Molecules 29 03812 g004
Figure 5. (a) O2-TPD spectrum and (b) EPR profiles of the catalysts.
Figure 5. (a) O2-TPD spectrum and (b) EPR profiles of the catalysts.
Molecules 29 03812 g005
Figure 6. (a) Catalyst screening. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg catalysts, 30 °C, 0.5 h. (b) Effect of temperature. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 3 h. (c) Effect of time. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 40 °C. (d) Effect of catalyst dosage. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 40 °C, 3 h. (e) Effect of H2O2 dosage. Conditions: 0.1 mol NMM, 20 mg 5%Zr-TiO2, 40 °C, 3 h. (f) Cycling stability of 5%Zr-TiO2.
Figure 6. (a) Catalyst screening. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg catalysts, 30 °C, 0.5 h. (b) Effect of temperature. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 3 h. (c) Effect of time. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 20 mg 5%Zr-TiO2, 40 °C. (d) Effect of catalyst dosage. Conditions: 0.1 mol NMM, 0.13 mol H2O2, 40 °C, 3 h. (e) Effect of H2O2 dosage. Conditions: 0.1 mol NMM, 20 mg 5%Zr-TiO2, 40 °C, 3 h. (f) Cycling stability of 5%Zr-TiO2.
Molecules 29 03812 g006
Scheme 1. The schematic diagram of the catalytic oxidation process.
Scheme 1. The schematic diagram of the catalytic oxidation process.
Molecules 29 03812 sch001
Table 1. The crystal size, surface area, and pore diameter of the catalysts.
Table 1. The crystal size, surface area, and pore diameter of the catalysts.
CatalystsCrystal Size
(nm)
Surface Area
(m2/g)
Pore Diameter
(nm)
TiO221.7166.1
1%Zr-TiO214.6495.7
3%Zr-TiO211.5725.1
5%Zr-TiO210.6855.0
7%Zr-TiO211.1755.4
10%Zr-TiO211.4745.5
Table 2. The surface compositions of all catalysts.
Table 2. The surface compositions of all catalysts.
Catalysts[Zr]/[Ti][O]/[Ti][C]/[Ti]Oads/(Oads + Olatt)%
TiO2--1.981.0012.6
1%Zr-TiO20.0071.990.7213.5
3%Zr-TiO20.0152.050.7815.0
5%Zr-TiO20.0212.060.7518.3
7%Zr-TiO20.0302.040.7715.4
10%Zr-TiO20.0452.150.7214.8
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

Li, Y.; Fang, Z.; Feng, L.; Liu, F.; Shi, Y.; Li, J.; Zhao, C. Study of Mesoporous Zr-TiO2 Catalyst with Rich Oxygen Vacancies for N-Methylmorpholine Oxidation to N-Methylmorpholine-N-oxide. Molecules 2024, 29, 3812. https://doi.org/10.3390/molecules29163812

AMA Style

Li Y, Fang Z, Feng L, Liu F, Shi Y, Li J, Zhao C. Study of Mesoporous Zr-TiO2 Catalyst with Rich Oxygen Vacancies for N-Methylmorpholine Oxidation to N-Methylmorpholine-N-oxide. Molecules. 2024; 29(16):3812. https://doi.org/10.3390/molecules29163812

Chicago/Turabian Style

Li, Yongwei, Zhihao Fang, Lijuan Feng, Fangfang Liu, Yucui Shi, Jiao Li, and Chao Zhao. 2024. "Study of Mesoporous Zr-TiO2 Catalyst with Rich Oxygen Vacancies for N-Methylmorpholine Oxidation to N-Methylmorpholine-N-oxide" Molecules 29, no. 16: 3812. https://doi.org/10.3390/molecules29163812

Article Metrics

Back to TopTop