Daojian
Tang‡
ac,
Lei
Wu‡
ac,
Liubo
Li
b,
Niankai
Fu
bc,
Chuncheng
Chen
ac,
Yuchao
Zhang
*ac and
Jincai
Zhao
ac
aKey Laboratory of Photochemistry, Beijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China. E-mail: yczhang@iccas.ac.cn
bKey Laboratory of Molecular Recognition and Function, Beijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China
cUniversity of Chinese Academy of Sciences, Beijing 100049, P. R. China
First published on 16th January 2024
Photo(electro)catalytic chlorine oxidation has emerged as a useful method for chemical transformation and environmental remediation. However, the reaction selectivity usually remains low due to the high activity and non-selectivity characteristics of free chlorine radicals. In this study, we report a photoelectrochemical (PEC) strategy for achieving controlled non-radical chlorine activation on hematite (α-Fe2O3) photoanodes. High selectivity (up to 99%) and faradaic efficiency (up to 90%) are achieved for the chlorination of a wide range of aromatic compounds and alkenes by using NaCl as the chlorine source, which is distinct from conventional TiO2 photoanodes. A comprehensive PEC study verifies a non-radical “Cl+” formation pathway, which is facilitated by the accumulation of surface-trapped holes on α-Fe2O3 surfaces. The new understanding of the non-radical Cl− activation by semiconductor photoelectrochemistry is expected to provide guidance for conducting selective chlorine atom transfer reactions.
Organochlorides are commonly found in many bioactive natural products or synthetic drugs and are widely employed for the construction of complex structures in organic synthesis.14–17 Conventional synthetic methods involve the use of massive amounts of explosive, corrosive, or toxic chlorine sources such as chlorine gas (Cl2), which raises concerns regarding potential environmental risks.16,18,19 The electrochemical technology for chlorine evolution has inspired an oxidative chlorination approach to generate reactive chloride species in situ, facilitating the green synthesis of high-value chemicals such as chlorohydrins, aryl chlorides, and epoxides.20–24 Alternatively, PEC-assisted oxidative chlorination is expected to be a clean and more energy-saving chemical process with reduced electrical power input.25 However, reports on the photoredox protocol for chlorine activation show challenges in PEC oxidative chlorination, where photocatalysts (e.g., TiO2,25–27 WO3,5,6etc.) with deep valence band edges are often used to drive the chloride (Cl−) oxidation (Fig. S1†). The photogenerated holes in the valence band (VB) of these photocatalysts are more positive than the 1-e− oxidation potential of Cl−, which forms chlorine radicals (Cl˙).25,27 Although Cl˙ facilitates the activation of C–H bonds,25,27,28 it can more easily trigger uncontrolled radical chain reactions, leading to polychlorination and radical coupling reactions.27,29 Thus, a compromise must be made between selectivity and conversion. As such, the development of a PEC approach to achieve more controlled activation of Cl− for C–Cl construction remains a substantial challenge.
Hematite (α-Fe2O3) is one of the most promising photoanode materials in PEC studies due to its abundance, non-toxicity, stability, and visible light absorption capabilities (Fig. S1†). Recently, α-Fe2O3 has attracted wide attention for selective oxygen transfer reactions.30,31 Under PEC conditions, long-lived surface-trapped holes have been demonstrated on α-Fe2O3,32,33 and multi-hole H2O oxidation catalysis is realized, in which the accumulation of multiple oxidizing equivalents (i.e., high-valent iron-oxo species, FeIVO and FeVO) has been proposed as the key to efficient H2O oxidation catalysis.34–36 It is further found that α-Fe2O3 serves as an efficient oxygen atom transfer (OAT) catalyst, achieving highly selective two-electron oxygenation of substrates such as thioethers and alkenes.30 The high-valent iron-oxo characteristics of surface-trapped holes on α-Fe2O3 is reminiscent of the high-valent iron-oxo species in halogenating enzymes, which are capable of controllably oxidizing Cl− to form “Cl+” species (e.g., HClO or its equivalents) through a two-electron oxidation process.15,37,38 This enables selective electrophilic chlorination for constructing C–Cl bonds and preventing side reactions caused by single-electron radical reactions under relatively mild reaction conditions. There has been a significant advancement in the spectroscopic, physicochemical, and mechanistic investigations of homogeneous Fe catalysts, but the chlorination reaction using heterogeneous Fe or Fe oxide catalysts is rarely reported.38–40
Inspired by this, we envision that α-Fe2O3 could serve as an efficient PEC catalyst for controlled non-radical activation of Cl− for C–Cl construction. In this work, the Cl− oxidation process on α-Fe2O3 is investigated. We show a direct “Cl+” species formation process on α-Fe2O3 surfaces, which is facilitated by the accumulation of surface-trapped holes. As a result, α-Fe2O3 exhibits a high selectivity (up to 99%) for the synthesis of aromatic chlorides and chlorohydrins. This is in sharp contrast to the low-selective chlorination on TiO2 photoanodes, which follows the free radical chlorination pathway. This work is anticipated to offer an in-depth understanding of selective Cl− activation by semiconductor photoelectrochemistry and provide guidance for the construction of efficient chlorine atom transfer reactions.
Intensity modulated photocurrent spectroscopy (IMPS) measurements were performed to compare interfacial charge-transfer kinetics between the oxidations of Cl− and H2O on the α-Fe2O3 photoanode. IMPS plots showed identical semi-circles in the high frequency domain (the fourth quadrant) but a decreased radius of the low-frequency circle in the first quadrant after the addition of Cl− (Fig. S4†). Simulation of the IMPS plots showed that the presence of Cl− increased the charge-transfer rate constant (kct) from 13.4 s−1 to 21.6 s−1 and the charge-transfer efficiency (ηct) from 53.5% to 77.0%, indicating that Cl− oxidation was much faster than H2O oxidation. The inhibition of H2O oxidation was confirmed by monitoring the generation of O2 by gas chromatography (GC, Fig. S5†) after 1 h of photoelectrolysis under AM 1.5G irradiation. As shown in Fig. 1b and Table S1,† the addition of Cl− resulted in a rapid decrease in the faradaic efficiency (FE) of the oxygen evolution reaction (OER) and almost complete suppression when 25–200 mM Cl− was present. To further confirm the high selectivity of Cl− oxidation, we employed the N,N-diethyl-p-phenylenediamine (DPD) colorimetric method to quantify the active chlorine concentrations under various Cl− solution concentrations (Fig. S6†).41 The results showed that after 10 min of photoelectrolysis, the FE of active chlorine generation was more than 90% when Cl− concentrations ranged from 25 to 200 mM, which was consistent with the results obtained from OER experiments. In addition, no active chlorination generation was detected under conditions of only electricity or light, suggesting that both light and electricity were necessary in our system for Cl− oxidation (Fig. S7†). Strikingly, α-Fe2O3 exhibited a high selectivity of Cl− oxidation even at a relatively low Cl− concentration of 50 mM, showing a significant advantage compared to other reported photocatalysts for Cl− oxidation (Table S2†). We also explored the Cl− oxidation performance in an aqueous solution. As shown in Fig. S8,† the selectivity of Cl− oxidation decreased in 0.1 M Cl− aqueous solution compared to that in a H2O–MeCN mixture solution (Fig. S6†), suggesting that the presence of an organic solvent helped to inhibit the competitive H2O oxidation reaction. Moreover, a further increase of Cl− concentration could improve the selectivity of Cl− oxidation. Under 0.3 M Cl−, the FE of Cl− oxidation could reach more than 90%, which still showed good selectivity for Cl− oxidation compared to other reported photocatalysts (Table S2†). These results indicated the high reactivity of α-Fe2O3 towards Cl− oxidation.
Previous research has demonstrated that the process of H2O oxidation on α-Fe2O3 occurs through surface-trapped holes rather than direct hole transfer from the valence band.42,43 In order to uncover the underlying relationship between the Cl− oxidation and the surface-trapped holes, electrochemical impedance spectroscopy (EIS) measurements were carried out under the same conditions to investigate the interface reaction of Cl− oxidation and H2O oxidation on α-Fe2O3. As shown in Fig. 1c, the Nyquist plots for both Cl− oxidation and H2O oxidation on the α-Fe2O3 photoanode consisted of two semicircles, and its Bode plots exhibited two peaks at different frequency regions (Fig. 1d and e). This is a typical feature of surface-state mediated charge transfer rather than direct hole transfer.3,43 The high-frequency (HF) semicircle in Nyquist plots should be assigned to the surface hole-trapping process and the low-frequency (LF) semicircle should be attributed to the transfer of those trapped holes to the electrolyte. In Fig. 1c, the HF semicircles were almost unchanged upon adding 10 mM Cl−, which indicated that Cl− had little effect on the generation of surface trapped holes. To further confirm this, potential-dependent experiments were conducted. For both H2O oxidation and Cl− oxidation, the potential-dependent experiments (Fig. 1d and e) showed the phase angle of the HF region (P1, which was assigned to the surface hole-trapping process) in Bode plots increased with applied potentials, suggesting the accumulation of surface-trapped holes. Regardless of the presence of Cl−, there was no distinction in the phase angle P1, demonstrating that Cl− oxidation and H2O oxidation shared similar surface states or active sites. This was consistent with the competitive oxidation process of Cl− and H2O (Fig. 1b and S6†). The trapped holes on these surface states were previously assigned to high-valent surface iron-oxo species, i.e., FeIVO and FeVO.43,44 On the other hand, compared to H2O oxidation, the radius of the LF semicircle in Cl− oxidation was significantly reduced (Fig. 1c). This demonstrated that the oxidation of Cl− mediated by surface-trapped holes was easier than the oxidation of H2O, which correlated well with the increasing photocurrent and the IMPS results.
To quantify the two associated capacitive elements, a well-established physical model consisting of the space-charge capacitance (Cbulk) and surface-state capacitance (Ctrap) was considered (see details in Fig. S9†).3,43–45 As shown in Fig. 1f, in 0.1 M NaClO4 solution, the fitted Ctrap increased with the applied bias at low potentials and reached a maximum at approximately 0.85 V due to the accumulation of surface-trapped holes. After that, Ctrap began to decrease because of the increased consumption rate of surface-trapped holes. In the presence of Cl−, the maximum value of Ctrap was smaller than that in the 0.1 M NaClO4 solution, indicating that the consumption of surface-trapped holes by Cl− oxidation were faster than that by H2O oxidation. To further prove that Cl− oxidation was mediated by the high-valent iron-oxo species, the behavior of surface-state capacitance (Ctrap) with Cl− concentrations was investigated (Fig. S10†). The Ctrap values decreased with the concentration of Cl−, supporting that Cl− oxidation was favorably mediated by the high-valent iron-oxo species. Therefore, we anticipated an efficient PEC chlorination of α-Fe2O3 through a controlled non-radical Cl− activation pathway similar to that of halogenating enzymes.
To investigate the performance of aromatic chlorination on α-Fe2O3, chronoamperometry was conducted at 1.0 V vs. Ag/AgCl under AM 1.5G irradiation for 2 h under the same conditions. Product analysis by high-performance liquid chromatography (HPLC) showed that 1a was converted into chlorinated products 2a and 3a (Fig. S13†). The FE for chlorinated products 2a and 3a on the α-Fe2O3 photoanode was found to be 90%, with a remarkable selectivity of 98% for monochlorinated product 2a (p/o = 1.6:1) (Fig. 2b). It has been reported that the TiO2 photoanodes drive Cl− oxidation to form ˙Cl through a single-electron oxidation mechanism.25 To confirm the uniqueness of α-Fe2O3 for efficient aromatic chlorination, the oxidative chlorination of 1a on a TiO2 photoanode was investigated and compared with that on α-Fe2O3. Structural characterization of TiO2 is shown in Fig. S14.† The FE for aromatic chlorination of 1a on TiO2 was only 50%, while the selectivity for 2a was 57% (Fig. 2b). In addition, the TiO2 photoanode showed a higher 3a/2a ratio for chlorinated products compared to that on the α-Fe2O3 photoanode, suggesting a low regioselectivity for aromatic mono-chlorination. This was consistent with the polychlorination occurring during the ˙Cl-involved chlorination reaction. Potential-dependent experiments on the α-Fe2O3 photoanode revealed that a high FE (>90%) was maintained within the potential range of 0.7 to 1.0 V (Fig. S15a†). By extending the electrolysis time to 2.5 h, a yield of 95% for 2a was achieved on the α-Fe2O3 photoanode (Fig. S15b†). We evaluated the stability of the α-Fe2O3 photoanode in batch reactions. Under high conversion of the substrate (>90%), a high selectivity of 98% was maintained for the 6th run under the same reaction conditions (Fig. 2c). The X-ray diffraction (XRD), X-ray absorption fine structure spectra (XAFS) and X-ray photoelectron spectroscopy (XPS) characterizations (Fig. S16 and 17†) displayed no changes for the used α-Fe2O3, demonstrating the stability of α-Fe2O3. Moreover, the presence of obvious Cl− residues was observed after photoelectrolysis, indicating a strong absorption of Cl− on Fe2O3 surfaces (Fig. S17c and d†). More importantly, the enhanced interfacial charge transfer of Cl− oxidation resulted in a higher solar H2 production (47 μmol h−1) compared to H2O oxidation (22 μmol h−1) as the oxidative half reaction (Fig. S18†), suggesting a higher utilization efficiency of photo-induced charges and a more effective storage of solar energy through the coupled electrophilic chlorination reaction with H2 production.
Considering the high Cl− oxidation performance across a wide range of Cl− concentrations (10–200 mM, Fig. 1b and S6†), the aromatic chlorination of 1a with different Cl− concentrations was also evaluated. At the same potential (1.0 V vs. Ag/AgCl), the yield rate of 2a increased with the higher concentration of Cl−, reaching 23 μmol cm−2 h−1, while the selectivity of 2a remained up to 90% (Fig. S15c and d†). This was consistent with the linear change in Cl− oxidation photocurrent (Fig. S3b†), indicating that the rate-determining step (RDS) of aromatic chlorination was the PEC Cl− oxidation rather than the subsequent chemical process. More importantly, under low Cl− concentration (10 mM Cl−, 1 equiv. substrate concentration), the chlorination activity on the α-Fe2O3 photoanode still showed a high FE of 75%, indicating the efficient Cl− oxidation and the rapid aromatic chlorination process on the α-Fe2O3 photoanode. The effective chlorination process on the α-Fe2O3 photoanode was also supported by inhibiting the further oxidation of active chlorine species to form ClO3− (Fig. S19†). During photoelectrolysis without 1a, the analysis of Cl− oxidation products using ion chromatography (IC) showed a gradual accumulation of ClO3− over a period of 2 h (Fig. S19a†). However, when photoelectrolysis was performed in the presence of 1a, ClO3− formation was completely inhibited (Fig. S19b†) and the chlorination of 1a showed a high current efficiency (FE 98%), indicating the rapid reaction between 1a and active chlorine species. This was further supported by DPD colorimetric tests,41 in which most of the active chlorine species derived from Cl− oxidation already participated in the aromatic chlorination and only a small portion of them was residual and detected on the α-Fe2O3 photoanode (Fig. 2d). In contrast, DPD colorimetric tests showed a significant accumulation of active chlorine species on the TiO2 photoanode (Fig. 2d), suggesting that most of the active chlorine species generated from Cl− oxidation on TiO2 were unable to participate in the subsequent aromatic chlorination and were thus detected by DPD tests. ˙Cl can lead to uncontrolled hydrogen abstraction and polymerization reactions, resulting in the low aromatic chlorination activity.18 These results suggested that PEC activation of Cl− on α-Fe2O3 and TiO2 photoanodes produced distinct active chlorine species.
To further confirm this, we explored the synthesis of chlorohydrin products (Fig. 2a, model reaction II), which act as essential intermediates in industrial-scale epoxidations. ˙Cl can initiate radical chain reactions or radical coupling reactions, resulting in the formation of dimeric or polymeric products during the chlorination of alkene.47,48 After electrolysis at 1.0 V vs. Ag/AgCl for 2 h, product analysis by HPLC showed that nearly 80% conversion of 1b with up to 85% selectivity of chlorohydrin 2b was achieved on the α-Fe2O3 photoanode (Fig. 2e and S20†). The calculated FE was close to 80%. In addition to producing chlorohydrin products of 2b, a minor quantity of dichloride products of 3b was also observed (Fig. 2e). In contrast, the chlorohydrin of 1b on the TiO2 photoanode only exhibited an FE of 23% and a selectivity of 21% for 2b. It also showed a much higher production of dichlorination products 3b (Fig. 2e). The gas chromatography-mass spectrometry (GC-MS, Fig. 2f) revealed the formation of high molecular mass polymers on the TiO2 photoanode, which were not observed on the α-Fe2O3 photoanode. These results indicated that ˙Cl is not formed and the selective electrophilic chlorination process occurs on α-Fe2O3.
a typical reaction conditions applied in this work: 0.1 M NaCl, 0.1 mmol substrate in 10 mL of CH3CN solution with 50% H2O (v/v = 1:1), PEC reactions were carried out for 2 h under AM 1.5G irradiation at 1.0 V vs. Ag/AgCl with a non-conditioned air atmosphere and pH adjustments. b reaction in 0.1 M NEt4BF4 with 16% H2O and 50 mM NaCl for 2 h. c reaction in 0.1 M NEt4BF4 with 16% H2O and 50 mM NaCl for 5 h. |
---|
In addition, the electrosynthesis of chlorohydrins was explored. As shown in Table 1, our PEC approach exhibited a high yield for the chlorohydrination of styrene and its derivatives (12–15) (Fig. S30–32†), thus demonstrating the universality of the PEC halogenation strategy on α-Fe2O3 photoanodes for conducting efficient chlorine atom transfer reactions.
Based on the kinetics of various mono-substituted benzene substrates in the aromatic chlorination reaction, the Hammett linear free energy relationship can be calculated. It showed a negative slope of −1.1 for para-chlorinated products on the α-Fe2O3 photoanode (Fig. 3a), indicating the positive charge buildup at the aromatic moiety of the substrate in the transition state. This confirmed the electrophilic chlorination pathway mediated by non-radical “Cl+” species on α-Fe2O3 photoanodes. This was further supported by toluene chlorination experiments (Fig. 3b). The “Cl+” species only triggers electrophilic aromatic substitution reactions, while highly reactive ˙Cl easily induces competitive sp3 C–H chlorination, resulting in low selectivity during the ˙Cl-mediated aromatic chlorination process (Fig. S33†). As shown in Fig. 3b, TiO2 generated few aromatic chlorination products (chlorotoluene, CT) in toluene chlorination experiments, but mainly produced the sp3 C–H chlorination product benzyl chloride (BC). In contrast, the α-Fe2O3 photoanode exhibited a high CT yield rate of 12.6 μmol cm−2 h−1 at 1.0 V vs. Ag/AgCl (Fig. 3b), suggesting that ˙Cl was the main active chlorine species on the TiO2, while the active chlorine species generated on the α-Fe2O3 were more likely to be “Cl+” species rather than ˙Cl. To further confirm this, electron paramagnetic resonance (EPR) measurements were carried out to detect possible radicals. 5,5-Dimethyl-1-pyrroline N-oxide (DMPO) was used as the spin-trapping reagent to capture the possible radicals that were generated in the photoelectrolysis. As shown in Fig. 3c, radical signals were detected only on the TiO2 photoanode. The three peaks at 3338, 3353, and 3368 G were attributed to DMPO-Cl˙, which was in situ produced by the reaction between DMPO and Cl˙ during the photoelectrolysis on TiO2 (red rhombi). The quartet peaks at 3332, 3345, 3360, and 3374 G with an intensity ratio of 1:2:2:1 corresponded to DMPO-OH˙ (yellow stars). The other six peaks indicated the formation of DMPO-OCl˙ (blue hearts).49,50 These EPR experiments supported the radical process on TiO2 and non-radical process on α-Fe2O3.
Cl2 is also one of the active chlorine species and can be formed by the self-coupling of Cl˙. To rule out the possibility that Cl2 is the main active chlorine species on α-Fe2O3, substrate generation/tip collection mode scanning electrochemical microscopy (SG/TC SECM) measurements were conducted (see more details in Fig. S35 and 36†) to directly monitor the formation of Cl2 on the photoanodes surfaces. As shown in Fig. 3d, Cl2 was used as the probe molecule to specifically detect Cl2 generated from the photoanode surface at an ultra-close distance (∼13 μm). With the increase in photocurrent (Isub) after an onset potential of −0.2 V vs. Ag/AgCl, the corresponding rise in Itip unequivocally confirmed the formation of Cl2 on TiO2 surfaces (Fig. 3e). In contrast, the production of Cl2 was significantly lower on α-Fe2O3 surfaces, especially when the applied potential exceeded 0.9 V vs. Ag/AgCl. For example, at 1.2 V vs. Ag/AgCl, the Isub for Cl− oxidation on α-Fe2O3 was significantly higher than that on TiO2. In contrast, the Itip for Cl2 production was more than 4 times lower on α-Fe2O3 compared to that on TiO2, demonstrating that Cl2 was not the main product on α-Fe2O3 surfaces. The residue amount of Cl2 on α-Fe2O3 may originate from the reaction between “Cl+” species and Cl−.
To further explore the nature of the Cl− activation process on α-Fe2O3, rate law analysis based on EIS measurements was performed to compare the hole transfer kinetics for Cl− oxidation between α-Fe2O3 and TiO2 photoanodes (see details in Fig. S9†). Cl− oxidation on α-Fe2O3 showed a second-order kinetics (Fig. 3f and S37, and Tables S4 and S5†). In contrast, Cl− oxidation on TiO2 exhibited a first-order kinetics. The second-order reaction kinetics indicated that two-hole transfer was involved in the RDS for the Cl− oxidation reaction on α-Fe2O3, indicating that “Cl+” species were directly formed from the reaction between the absorbed Cl− and the surface-trapped holes (i.e., high-valent iron-oxo species). To support this conclusion, HClO was purposely used as a model “Cl+” reagent for the chlorination of 1a. As shown in Fig. S38,† its product distribution was similar to that under the PEC reaction conditions, implying that “Cl+” species (e.g., HClO or its equivalents) were the actual active chlorine species during the PEC Cl− oxidation on α-Fe2O3 for the subsequent electrophilic chlorination. In contrast, the Cl− oxidation on TiO2 was dominated by transfer of a single surface-trapped hole, which is consistent with the free radical characteristics of TiO2.
Moreover, we also successfully constructed a self-powered PEC system for the electrophilic chlorination reaction. In this system, a solar panel (∼3 V) was used to provide a constant bias (Fig. 4a). As shown in Fig. 4b, this self-powered PEC system successfully achieved aromatic chlorination of acetanilide, showing a promising approach to harness solar energy for synthesizing valuable organic halides.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3sc06337b |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2024 |