Fluorine-functionalized zirconium-organic cages for efficient photocatalytic oxidation of thioanisole

Jia-Ru Li Ning Li Li-Ling He Jun He

Citation:  Jia-Ru Li, Ning Li, Li-Ling He, Jun He. Fluorine-functionalized zirconium-organic cages for efficient photocatalytic oxidation of thioanisole[J]. Chinese Chemical Letters, 2025, 36(1): 109934. doi: 10.1016/j.cclet.2024.109934 shu

Fluorine-functionalized zirconium-organic cages for efficient photocatalytic oxidation of thioanisole

English

  • Metal-organic cages (MOCs) are individual supramolecular units created from metal nodes and organic connectors. Despite their assembly and structure being similar to metal-organic frameworks (MOFs), the MOCs, structured at the nanoscale level, exhibit unique attributes in terms of size, topology, solubility, characterization, and stability [1]. So far, a variety of MOCs have been documented, including those derived from rhodium (Rh), iron (Fe), cobalt (Co), titanium (Ti), and zirconium (Zr) [2-6]. The high bond energies of metal-oxygen bonds impart considerable structural stability to these MOCs. In particular, the Zr-O bond, with its energy of 766.1 ± 10.6 kJ/mol, results in Zr-MOCs exhibiting greater stability compared to MOCs based on other metals. Nevertheless, there have been few reported instances of Zr-MOCs, with research into their application performances also being greatly limited [7,8]. During the synthesis of Zr-MOCs, the susceptibility of Zr to oxidation poses a challenge. Zr being a reactive transition metal, readily reacts with oxygen to form oxides, especially under conditions of elevated temperature, humidity, or exposure to air. This tendency contributes to the scarcity of reported Zr-MOCs structures, thereby limiting their applications in performance.

    To expand the performance research of Zr-MOCs, functionalizing modifications of the ligands of MOCs can be employed. Numerous studies have shown that incorporating functional groups into the organic ligands of MOCs can enhance their applications in various aspects [1,9-11]. Introducing Lewis basic functional groups (such as hydroxyl, amino, carboxyl, cyano, carbonyl, and fluorine) can enhance the selectivity and catalytic efficiency of specific organic reactions. Fluorine-functionalized MOCs exhibit characteristics such as fluorophilicity, electrostatic interactions, and anion-π stacking. Moreover, due to the high electronegativity of fluorine, the C-F bond possesses high bond energy (110 kJ/mol), leading to the high chemical stability of C-F bonds [12-17]. Furthermore, the high electronegativity of fluorine results in greater electron density toward the fluorine side in covalent bonds. This induces a shift in the electron energy levels within the molecular orbitals (MOs), elevating energy levels for atoms connected to fluorine and lowering energy levels for fluorine atoms. This facilitates redox reactions, making MOCs more widely applicable. Among the reported Zr-MOCs, the most classic are ZrT-1, 2, 3, 4 reported by Yuan in 2013 [18]. Over the past decade, various research groups have conducted numerous functionalization modifications on the cages reported by Yuan [18]. For example, Su’s group reported amino modifications on ligands, enabling their application in photocatalytic hydrogen evolution reactions [19]. However, their application in photocatalytic organic reactions is relatively limited.

    Building upon the achievements of Yuan in 2013 [18], where the classical ZrT-3 was employed, we fluorinated the organic ligands to synthesize H2bpdc-8F as a fluorine-rich linker (Fig. 1). Using solvent-thermal synthesis, we then combined H2bpdc-8F with Zirconocene dichloride (Cp2ZrCl2) to produce a tetrahedral-shaped Zr-MOCs: ZrT-3-F, with a molecular formula of {[Cp3Zr3μ3-O(μ2-OH)3]4(bpdc8F)6}·4Cl·nS (where S = noncoordinated solvent molecule). It is noteworthy that the fluorinated cage exhibited higher stability and enhanced light absorption capabilities. Utilizing these advancements, we applied the fluorinated cage for the first time in the oxidation of benzyl sulfide to the sulfoxide reaction. In photocatalytic reaction experiments, we observed that the fluorinated cage demonstrated a selectivity and conversion rate of over 95% for sulfoxide, significantly surpassing the performance of the non-fluorinated cage.

    Figure 1

    Figure 1.  Contrast sample: ZrT-3 (left), ZrT-3-F (right), bridging ligands and trinuclear zirconium clusters.

    The compound ZrT-3-F is produced via a solvent-thermal procedure, combining Cp2ZrCl2 and H2bpdc-8F in a mixed solvent environment of dimethylacetamide (DMA), trichloromethane (CHCl3), and H2O. The exact synthesis route can be found in the supplementary materials provided. Utilizing X-ray single crystal diffraction analysis, we determined that ZrT-3-F crystallizes in the cubic Fm-3 m space group. The compound forms a distinctive V4E6 type cage, as previously documented by Yuan. The structure of ZrT-3-F consists of four secondary building units (SBUs), namely Cp3Zr3μ3-O(μ2-OH)3, and six H2bpdc-8F linkers. The Cp3Zr3μ3-O(μ2-OH)3 unit acts as the vertex (V), while H2bpdc-8F functions as the edges (E) of the tetrahedron. Due to the high symmetry of ZrT-3-F, the linking ligands exhibit significant disorder. The interspace between two SBUs measures 15.387 Å, resulting in an approximate internal volume of the cage of 139.2 Å3, as derived from calculations.

    To evaluate the high photocatalytic activity potential of ZrT-3-F, a thorough investigation into its optical and electronic properties was conducted. Additionally, a comparative analysis between ZrT-3-F and ZrT-3 was carried out to elucidate key disparities in band structure and charge carrier behavior. UV/visible diffuse reflectance spectroscopy (DRS) measurements indicated that ZrT-3-F exhibited nearly full-spectrum absorbance, including visible light absorption between 400 nm and 480 nm, unlike ZrT-3 which primarily absorbed in the UV region (Fig. 2a). Bandgap energies were computed using the Kubelka-Munk function equation: αhv = A(hvEg)1/2. The direct bandgap energies of ZrT-3-F and ZrT-3 were estimated to be 2.41 and 3.32 eV, respectively, based on transformed differential spectra (Fig. 2b). Electronic band structures were determined by estimating conduction band (CB) potentials via Mott-Schottky (M-S) plots at frequencies of 1000, 1500, and 2000 Hz in the dark. Both ZrT-3-F and ZrT-3 M-S plots exhibited positive slopes across all frequencies, consistent with n-type semiconductor characteristics (Figs. S9 and S10 in Supporting information) [20]. Flat band potentials (VFB) were extrapolated to be –0.757 and –0.886 V vs. Ag/AgCl for ZrT-3-F and ZrT-3, respectively. Consequently, the CB potentials for ZrT-3-F and ZrT-3 were –0.557 and –0.686 V, respectively. Using the formula ECB = EVBEg [21], VB positions for ZrT-3-F and ZrT-3 were calculated as 1.853 and 2.634 V, respectively, corresponding to the band structure arrangement (Fig. 2c). Furthermore, photocurrent response tests revealed a substantial increase in current under light irradiation compared to dark conditions. Notably, the photocurrent of ZrT-3-F consistently exceeded that of ZrT-3, irrespective of light exposure (Fig. 2d).

    Figure 2

    Figure 2.  (a) UV–vis DRS spectra of ZrT-3-F (red) and ZrT-3 (black). (b) Band gaps of ZrT-3-F (red) and ZrT-3 (black). (c) Electronic band structure. (d) Photocurrent responses of ZrT-3-F (red) and ZrT-3 (black).

    The process of converting sulfides into sulfoxides via aerobic oxidation finds extensive usage in sectors such as fossil fuel desulfurization, medicinal synthesis, and wastewater management [22,23]. In conventional synthesis methods, the elevated oxidation potentials of sulfides may induce over-oxidation reactions, generating undesired by-products [24-26]. Consequently, an efficacious photocatalyst is required for the successful aerobic oxidation of sulfides to sulfoxides [27,28].

    An investigation on the photocatalytic performance of ZrT-3-F was carried out, utilizing the aerobic oxidation of sulfides to sulfoxides reaction. To determine the type of reaction products, a gas chromatography-mass spectrometry (GC–MS) was employed and quantification was conducted using a 50-bit ALS gas chromatograph. Since the main light absorption of ZrT-3-F lies in the ultraviolet region, UV light was adopted as the light source for the catalytic reaction. Initially, the choice of the best solvent was required, using oxygen as the oxidant. The selected solvents: N,N-Dimethylformamide (DMF), methanol (CH3OH), isopropanol (IPA), and acetonitrile (CH3CN) were put under consideration. The conversion and selectivity of CH3OH and CH3CN are the highest with the catalyst being homogeneous and heterogeneous in CH3OH and CH3CN respectively. From the standpoint of recyclability, the heterogeneous catalyst is more appropriate and hence CH3CN was selected as the reaction solvent. Subsequently, it was necessary to elect the right oxidant: even though O2, hydrogen peroxide (H2O2), and tert-butyl hydroperoxide (TBHP) all achieved conversion rates up to 99%, O2 proved to be the most optimal oxidant regarding product selectivity. Upon conducting experiments with varying masses of the catalyst, the results indicate that the optimal catalyst loading is 10 mg.

    The optimal experimental conditions are as follows: 0.1 mmol of substrate, 3 mL of CH3CN, and 10 mg of catalyst are added into a 10-mL quartz tube. The photocatalytic reaction is then performed for 6 h under 1 atm of oxygen and ultraviolet light irradiation. When ZrT-3-F is used as the catalyst, the conversion rate reaches up to 99.51% and selectivity up to 95.16%. In contrast, when ZrT-3 is used as the catalyst in the experiment, the conversion rate of benzyl sulfide is only 21.44%, and the selectivity is only 66.7%. When the organic ligand H2bpdc-8F is used, the conversion rate is 52.67%, and the selectivity is 96.67%. When Cp2ZrCl2 is used as the catalyst, the conversion rate is 51.48%, and the selectivity is 58.29%. In summary, in this experimental system, the catalyst ZrT-3-F has the highest value (Table 1).

    Table 1

    Table 1.  Photocatalytic oxidation of thioanisole catalyzed by the ZrT-3-F.
    DownLoad: CSV

    Next, the wide-ranging suitability of ZrT-3-F as a heterogeneous photocatalyst was put under scrutiny, assessing the transformation rates and selectivity amongst various sulfides. As shown in the table, all the chosen sulfides demonstrated over 90% selectivity and conversion to sulfoxides. Table 2 shows a marked drop in the conversion performance when the para-position on the sulfur is electron-absorbing, confirming that electron-donating para-substituents are more conducive to the photocatalytic oxidation of thioanisole to its corresponding sulfoxide under aerobic conditions.

    Table 2

    Table 2.  Photocatalytic selective oxidation of sulfides to sulfoxide over ZrT-3-F.
    DownLoad: CSV

    The photostability of ZrT-3-F was evaluated through several centrifugations processed and washed with CH3CN. Upon conducting three repeated experiments with ZrT-3-F (Fig. 3a), it was found that the reactive capacity of the catalyst did not decrease significantly. The catalyst harvested after the cycle was put through Fourier transform infrared (FT-IR) spectroscopy (Fig. S16 in Supporting information), with the spectral results demonstrating negligible differences between the pre-reaction and post-reaction spectrums. This indicates that the structure remained unaffected. According to these findings, ZrT-3-F demonstrates exceptional stability as a heterogeneous photocatalyst.

    Figure 3

    Figure 3.  (a) Stability of the ZrT-3-F photocatalyst for the selective aerobic oxidation of thioanisole under ultraviolet light. (b) Effect of scavengers on the photocatalytic oxidation of thioanisole. Reaction conditions: thioanisole (0.1 mmol), ZrT-3-F (10 mg), CH3CN (3.0 mL), O2 (1 atm), 300 W Xe lamp (λ = 200–400 nm), room temperature, 6.0 h. EPR detection of in situ formed (c) 1O2 using TEMP as a spin-trapping agent and (d) O2 radical using DMPO as radical trapping agent.

    To delve into the fundamental reaction mechanism responsible for the photocatalytic oxidation of thioanisole into sulfoxide under aerobic conditions, a sequence of radical quenching experiments was performed (Fig. 3b). Utilizing 2,6-di-tert-butyl-p-cresol (BHT) as a radical quenching agent, the production of reaction outputs was inhibited to a significant extent. This finding denotes the involvement of free radicals in this reaction. The subsequent introduction of copper sulfate (CuSO4), acting as an electron inhibitor, caused a noticeable reduction in catalytic conversion efficiency. The implementation of potassium iodide (KI) as a hole inhibitor induced a state wherein the substrate underwent nearly no conversion, attesting to the occurrence of a photo-induced electron-hole (e-h+) separation process in the reaction [29,30]. To detect the presence of reactive oxygen species (ROS) generated during the reaction, tert-butyl alcohol was added to quench any potential hydroxyl radicals (OH). The lack of any conspicuous effect on conversion rates post its addition provided the evidence that OH radicals are not produced within this reaction framework. When p-benzoquinone and 2,2,6,6-tetramethylpiperidin-1-oxyl (TEMPO) were incorporated for trapping superoxide anion radical (O2) and singlet oxygen (1O2) respectively, a sharp decline in conversion rates was observed. This observation conclusively established that the ROS involved in this reaction are O2 and 1O2 [30,31].

    To further validate the existence of ROS in the reaction, electron paramagnetic resonance (EPR) experiments were implemented. The spin trap agent selected for the O2 was 5,5-dimethyl-1-pyrroline N-oxide (DMPO), while 2,2,6,6-tetramethylpiperidine (TEMP) was utilized for trapping singlet oxygen 1O2. As depicted in the accompanying Figs. 3c and d, both O2 and 1O2 demonstrated a significant response under photonic exposure, while no noticeable response was registered in dark conditions.

    In conclusion, the conceivable mechanism for the photolytic conversion of phenyl methyl sulfide to sulfoxide (Fig. 4) involves UV light irradiation on ZrT-3-F, instigating the production of photo-generated electrons and holes. These photo-generated electrons swiftly disperse towards the surface of the catalyst and into the conduits where O2 is reduced by the photoelectrons, resulting in O2. Neutral exciton-forming entities possessing robust Coulomb interactions undergo a transition from the singlet excited state (S1) to the low-energy triplet state (T1) via inter-system crossing (ISC) [32]. In this scenario, the energy is ceded to the attached O2 molecules, leading to the production of 1O2 and phosphorescence. Upon acquisition of a proton, the sulfide transforms into the free radical cation CH3SC6H5H+. O2 and CH3SC6H5H+ then initiate a nucleophilic reaction [33,34]. Concurrently, 1O2 instigates an electrophilic reaction with the sulfide to create a sulfur peroxide intermediate. The reaction of the sulfide with the sulfur peroxide intermediate paves the way for the synthesis of sulfoxides. Notably, the end product of both mechanisms is sulfoxide [35].

    Figure 4

    Figure 4.  Possible photocatalytic mechanisms.

    Fluorination of the classical Zr-MOCs, such as ZrT-3, results in enhanced light absorption compared to the non-fluorinated cages. The fluorinated ZrT-3 exhibits a stronger photocurrent response, thereby presenting advantages in photocatalytic reactions. Therefore, we employed ZrT-3-F for the oxidation of benzyl sulfide to sulfoxide. In photocatalytic reactions, ZrT-3-F serves as a heterogeneous catalyst under aerobic conditions, and achieves conversion rates and selectivity in oxidizing benzyl sulfide to sulfoxide. The conversion rate and selectivity are significantly higher compared to the performance of ZrT-3 in catalytic reactions, demonstrating the excellent versatility and chemical stability of the catalyst throughout the reaction process. Importantly, this work demonstrates the applicability of fluorinated Zr-MOCs in organic catalysis.

    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.

    Jia-Ru Li: Data curation, Investigation, Methodology, Writing – original draft, Writing – review & editing. Ning Li: Funding acquisition, Supervision, Writing – review & editing. Li-Ling He: Formal analysis, Validation. Jun He: Funding acquisition, Resources, Supervision.

    This work is supported by National Natural Science Foundation of China (Nos. 22201046, 22371054), Local Innovation Research Team Project of Guangdong Pearl River Talent Plan (No. 2017BT01Z032).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109934.


    1. [1]

      E.M. El-Sayed, Y.D. Yuan, D. Zhao, et al., Acc. Chem. Res. 55 (2022) 1546–1560. doi: 10.1021/acs.accounts.1c00654

    2. [2]

      S. Furukawa, N. Horike, M. Kondo, et al., Inorg. Chem. 55 (2016) 10843–10846. doi: 10.1021/acs.inorgchem.6b02091

    3. [3]

      I. Papadaki, C.D. Malliakas, T. Bakas, et al., Inorg. Chem. 48 (2009) 9968–9970. doi: 10.1021/ic901145d

    4. [4]

      S.T. Wang, X.H. Gao, X.X. Hang, et al., J. Am. Chem. Soc. 138 (2016) 16236–16239. doi: 10.1021/jacs.6b11218

    5. [5]

      Y.P. He, L.B. Yuan, G.H. Chen, et al., J. Am. Chem. Soc. 139 (2017) 16845–16851. doi: 10.1021/jacs.7b09463

    6. [6]

      Z.F. Ju, G.L. Liu, Y.S. Chen, et al., Chem. Eur. J. 23 (2017) 4774–4777. doi: 10.1002/chem.201700798

    7. [7]

      M.G. Sullivan, H.K. Welgama, M.R. Crawley, et al., Chem. Mat. 36 (2024) 567–574. doi: 10.1021/acs.chemmater.3c02775

    8. [8]

      B.X. Jiang, H. Wang, Y.T. Zhang, et al., Polyhedron 243 (2023) 116569. doi: 10.1016/j.poly.2023.116569

    9. [9]

      H.Y. Lin, Z.F. Xiao, K.N. Le, et al., Angew. Chem. Int. Ed. 61 (2022) e202214055. doi: 10.1002/anie.202214055

    10. [10]

      J.J. Jiao, C.X. Tan, Z.J. Li, et al., J. Am. Chem. Soc. 140 (2018) 2251–2259. doi: 10.1021/jacs.7b11679

    11. [11]

      W.H. Xing, H.Y. Li, X.Y. Dong, et al., J. Mater. Chem. A 6 (2018) 7724–7730. doi: 10.1039/C8TA00858B

    12. [12]

      S. Lee, H. Jeong, D. Nam, et al., Chem. Soc. Rev. 50 (2021) 528–555. doi: 10.1039/D0CS00443J

    13. [13]

      Q. Tian, S. Chen, M. Shi, et al., Sens. Actuator. B: Chem. 404 (2024) 135309. doi: 10.1016/j.snb.2024.135309

    14. [14]

      H.T. Chen, T. Zhang, S.R. Liu, et al., Inorg. Chem. 61 (2022) 11949–11958. doi: 10.1021/acs.inorgchem.2c01686

    15. [15]

      Z.A. Qiao, S.S. Brown, J. Adcock, et al., Angew. Chem. Int. Ed. 51 (2012) 2888–2893. doi: 10.1002/anie.201107812

    16. [16]

      D.K. Dogutan, R. McGuire, D.G. Nocera, J. Am. Chem. Soc. 133 (2011) 9178–9180. doi: 10.1021/ja202138m

    17. [17]

      S.S. Wang, L. Wang, M. Dakovic, et al., ACS Catal. 2 (2012) 230–237. doi: 10.1021/cs200501n

    18. [18]

      G.L. Liu, Z.F. Ju, D.Q. Yuan, et al., Inorg. Chem. 52 (2013) 13815–13817. doi: 10.1021/ic402428m

    19. [19]

      M. Sun, Q.Q. Wang, C. Qin, et al., Chem. Eur. J. 25 (2019) 2824–2830. doi: 10.1002/chem.201805663

    20. [20]

      H.F. Ye, R. Shi, X. Yang, Y. Chen, et al., Appl. Catal. B: Environ. 233 (2018) 70–79. doi: 10.1016/j.apcatb.2018.03.060

    21. [21]

      X. Li, Y. Qi, G.Z. Yue, et al., Green Chem. 21 (2019) 649–657. doi: 10.1039/C8GC03295E

    22. [22]

      Y.M. Li, M. Wang, X.F. Jiang, ACS Catal. 7 (2017) 7587–7592. doi: 10.1021/acscatal.7b02735

    23. [23]

      A.K. Clarke, A. Parkin, R.J.K. Taylor, et al., ACS Catal. 10 (2020) 5814–5820. doi: 10.1021/acscatal.0c00690

    24. [24]

      Q. Li, X.W. Lan, G.Y. An, et al., ACS Catal. 10 (2020) 6664–6675. doi: 10.1021/acscatal.0c00290

    25. [25]

      X.Y. Wang, M.H. Liu, Y.X. Liu, et al., J. Am. Chem. Soc. 145 (2023) 26900–26907. doi: 10.1021/jacs.3c09699

    26. [26]

      X. Liang, Z.F. Guo, H.X. Wei, et al., Chem. Commun. 54 (2018) 13002–13005. doi: 10.1039/C8CC07585A

    27. [27]

      M. Forchetta, F. Sabuzi, L. Stella, et al., J. Org. Chem. 87 (2022) 14016–14025. doi: 10.1021/acs.joc.2c01648

    28. [28]

      X.N. Zou, D.S. Zhang, T.X. Luan, et al., ACS Appl. Mater. Interfaces 13 (2021) 20137–20144. doi: 10.1021/acsami.1c03083

    29. [29]

      H.X. Wei, Z.F. Guo, X. Liang, et al., ACS Appl. Mater. Interfaces 11 (2019) 3016–3023. doi: 10.1021/acsami.8b18206

    30. [30]

      C.L. Su, R. Tandiana, B.B. Tian, J. Su, K.P. Loh, et al., ACS Catal. 6 (2016) 3594–3599. doi: 10.1021/acscatal.6b00443

    31. [31]

      X.Y. Gu, X. Li, Y.H. Chai, et al., Green Chem 15 (2013) 357–361. doi: 10.1039/c2gc36683e

    32. [32]

      Q.Q. Li, P.H. Pan, H. Liu, et al., Inorg. Chem. 62 (2023) 17182–17190. doi: 10.1021/acs.inorgchem.3c02212

    33. [33]

      J. Jiang, R.C. Luo, X.T. Zhou, et al., Adv. Synth. Catal. 360 (2018) 4402–4411. doi: 10.1002/adsc.201800730

    34. [34]

      H.D. She, H. Zhou, L.S. Li, et al., ACS Sustain. Chem. Eng. 6 (2018) 11939–11948. doi: 10.1021/acssuschemeng.8b02217

    35. [35]

      Q.S. Zhu, H.Y. An, T.Q. Xu, et al., Appl. Catal. A: Gen. 662 (2023) 119283. doi: 10.1016/j.apcata.2023.119283

  • Figure 1  Contrast sample: ZrT-3 (left), ZrT-3-F (right), bridging ligands and trinuclear zirconium clusters.

    Figure 2  (a) UV–vis DRS spectra of ZrT-3-F (red) and ZrT-3 (black). (b) Band gaps of ZrT-3-F (red) and ZrT-3 (black). (c) Electronic band structure. (d) Photocurrent responses of ZrT-3-F (red) and ZrT-3 (black).

    Figure 3  (a) Stability of the ZrT-3-F photocatalyst for the selective aerobic oxidation of thioanisole under ultraviolet light. (b) Effect of scavengers on the photocatalytic oxidation of thioanisole. Reaction conditions: thioanisole (0.1 mmol), ZrT-3-F (10 mg), CH3CN (3.0 mL), O2 (1 atm), 300 W Xe lamp (λ = 200–400 nm), room temperature, 6.0 h. EPR detection of in situ formed (c) 1O2 using TEMP as a spin-trapping agent and (d) O2 radical using DMPO as radical trapping agent.

    Figure 4  Possible photocatalytic mechanisms.

    Table 1.  Photocatalytic oxidation of thioanisole catalyzed by the ZrT-3-F.

    下载: 导出CSV

    Table 2.  Photocatalytic selective oxidation of sulfides to sulfoxide over ZrT-3-F.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  74
  • HTML全文浏览量:  3
文章相关
  • 发布日期:  2025-01-15
  • 收稿日期:  2024-03-13
  • 接受日期:  2024-04-28
  • 修回日期:  2024-04-04
  • 网络出版日期:  2024-11-06
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章