Continuous photocatalytic preparation of hydrogen peroxide with anthraquinone photosensitizers

Zongyi Yu a, Shichang Li a, Yufeng Wu a, Cunfei Ma a, Jianing Li a, Liyuan Duan a, Zunchao Liu a, Huinan Sun a, Guofeng Zhao a, Yue Lu a, Qilei Liu *a, Qingwei Meng *ab and Jingnan Zhao *a
aState Key Laboratory of Fine Chemicals, Frontiers Science Center for Smart Materials Oriented Chemical Engineering, Department of Pharmacy, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China. E-mail: Mengqw@dlut.edu.cn
bNingbo Institute of Dalian University of Technology, Dalian University of Technology, Ningbo 315000, China

Received 15th May 2024 , Accepted 28th June 2024

First published on 11th July 2024


Abstract

Hydrogen peroxide (H2O2) is a versatile oxidant, and its eco-friendly synthesis via visible light and air oxygen is crucial. Anthraquinone compounds were employed as homogeneous photocatalysts for alcohol oxidation, yielding H2O2 under visible light. Instead of using traditional catalyst library screening, a model based on density functional theory was proposed to structurally optimize anthraquinone photocatalysts, resulting in the discovery of a novel catalyst significantly boosting H2O2 generation (up to 621.2 mM h−1 intermittently). Continuous flow reactor design improvements can enhance chemical transformation efficiency, as demonstrated by continuous H2O2 production using 2,6-di-tert-butyl anthraquinone. When isopropanol was used as the hydrogen atom donor, H2O2 production rates reached 3950.6 mM h−1 in a 7-minute reaction, with a space–time yield of 7.90 mol (L h)−1, showcasing promising advancements in green H2O2 synthesis.


Introduction

Hydrogen peroxide, a versatile oxidant with robust oxidative properties spanning the entire pH range,1 is employed in the oxidative preparation processes of various crucial chemical products, including fatty aldehydes, aromatic ketones, cyclohexanone oxime, and epoxy propane.2,3 Additionally, its oxidation reactions ultimately yield water, making it one of the cleanest and most environmentally friendly oxidants.4 Currently, the primary method for large-scale production of hydrogen peroxide (H2O2) is through the AQ (anthraquinone) process: anthraquinone reduction, hydrogenation, H2O2 extraction, and anthraquinone solution circulation.5 The commonly used anthraquinone catalyst is 2-ethylanthraquinone (EAQ), loaded at 120–140 g L−1 in the working solution. For reduction, H2 is used with Pd/C catalysts, achieving 9.0–12.0 g L−1 hydrogenation efficiency (HE).6 The hydrogenated anthraquinone catalyst, acting as a hydrogen carrier, enters a countercurrent oxidation tower for reduction with air oxygen, yielding H2O2. Water is then added for H2O2 extraction from the solution, which is recycled for subsequent reaction cycles.

However, the above-mentioned process has several issues. The high loading of 2-ethylanthraquinone catalysts, potential deactivation of Pd/C and, along with the complexity of working fluid composition, high equipment investment, and significant discharge of waste, make this process hazardous, complex, and costly. There is a need to develop an efficient, safe, and environmentally sustainable method for H2O2 production.5

The production of hydrogen peroxide (H2O2) through photocatalysis stands as a sustainable process, leveraging water and oxygen as primary source materials and solar energy as its driving force. Recently, photocatalytic H2O2 production has become a significant area of research.7 A wide range of photocatalytic strategies has been explored, mainly categorized into heterogeneous and homogeneous systems. In heterogeneous strategies for H2O2 production using water or organic sacrificial agents, the focus is on various inorganic semiconductors, such as g-C3N4,8–13 TiO2,14 ZnO,15 BiVO4[thin space (1/6-em)]16 and other doped semiconductor materials.17–21 Additionally, organic materials like polymers,21–25 COFs,26–29 MOFs30 and supramolecular photocatalyst.31–33 After modifications and improvements like nanostructure design, co-doping, metal loading, surface modification, and organic photosensitizer introduction, the highest H2O2 production yield achieved in heterogeneous photocatalytic strategies is 1.7 mM h−1 (approx. 1.7 mmol g−1 h−1). This was with Co(II)-doped TiO2 under 420 nm light.

Homogeneous photocatalytic strategies, typified by anthraquinone compounds like EAQ, have showcased superior catalytic efficiency in H2O2 production. Ren34 and team initially demonstrated H2O2 production using the industrial anthraquinone process's working solution, yielding 43.2 mM h−1. Chen35et al. utilized ethanol as a hydrogen donor, achieving 54.0 mM h−1 under simulated sunlight. Vibbert36et al. used 2-tert-butylanthraquinone (BAQ) and toluene, achieving 63.1 mM h−1 under 395 nm light. In our previous work,37 we achieved 323.2 mM h−1 using EAQ and isopropanol. While EAQ is commonly used, its susceptibility to oxidation in photocatalytic processes calls for novel, stable, and efficient photocatalysts for H2O2 photoproduction (Scheme 1).


image file: d4gc02385d-s1.tif
Scheme 1 Overview detailing the intelligent photocatalytic continuous production of H2O2.

To reveal the structure–performance relationships of anthraquinone catalysts and address the issue of low photocatalytic efficiency, a density functional theory (DFT)-based surrogate model was developed to identify novel anthraquinone photocatalysts with high catalytic performance and reduce the synthesis workload. The AQ catalysts were characterized with the electrostatic potential (ESP) on molecular surface and the dipole moment, which were selected as critical descriptors and utilized to establish the surrogate model for predictions of the production of hydrogen peroxide in photocatalysis using the least squares regression method. The use of the surrogate model yielded both qualitative understanding of the catalytic rules and quantitative identification of several potential AQ catalysts. These catalysts were subsequently synthesized and their enhanced performance in producing H2O2 was confirmed through experimentally validation.

Given that industrial H2O2 applications use aqueous solutions with varied mass fractions, this approach reduces H2O2 mass transfer efficiency in organic phases, weakening oxidation efficiency. Introducing the aqueous phase complicates post-processing in organic reactions. This work aims to establish a novel strategy and apparatus for continuous, on-demand high-concentration H2O2 organic phase solutions. Tailored to reactants and reaction needs, this approach enables real-time, in situ production of target mass fraction H2O2 solutions, ensuring an environmentally friendly, efficient, controllable, and inherently safe process.

Results and discussion

Establishment of catalyst design model

First of all, the DFT-based electrostatic potential of molecular surface was performed on the existing anthraquinone-based catalysts in the laboratory. The molecular simulation results were utilized to analyze and predict the photocatalytic performance of each catalyst in the production of H2O2.

Linear regression analysis conducted on five anthraquinone-based catalysts—2-ethylanthraquinone (EAQ), 2-tert-butylanthraquinone (BAQ), 2-bromoanthraquinone (BrAQ), 2-chloroanthraquinone (CAQ), and 9-thioxanthenone (TX)—yielded a high correlation (R2 = 0.9955) between the dipole moment (DM) descriptor computed by DFT and the measured rates of H2O2 production in photocatalysis42–46 (eqn (1), Fig. 1B and C). The regression results demonstrated that the dipole moment47–49 of catalysts within a certain range is closely associated with the photocatalytic performance of anthraquinone-based catalysts for H2O2 production. Accordingly, it was hypothesized that designing a novel catalyst and forecasting its photocatalytic performance for H2O2 generation can be realized by calculating the dipole moment via DFT calculations and applying eqn (1). Further experimental work confirmed this methodology using 2-methylanthraquinone (MAQ), which possessed a dipole moment of 0.6496 and was predicted to have an H2O2 yield of 535.4 mM h−1. This prediction had a relative error of only 1.81% from the experimentally measured yield of 525.7 mM h−1 (Fig. 1D and E), highlighting the superior catalytic performance of MAQ over BAQ. Additionally, the prediction for 2-tert-amylanthraquinone (TAAQ) (a long-chain alkane-substituted derivative) showed an H2O2 production of 451.4 mM h−1 predicted by the dipole moment of 0.9365, deviated by only 2.39% from the actual yield of 440.6 mM h−1 (Fig. 1D and E). Considering catalyst deactivation risks due to oxidation and minimizing active C–H bonds, 2,6-di-tert-butylanthraquinone (DBAQ) was designed and synthesized with a dipole moment of 0.0007, predicting an H2O2 yield of 717.0 mM h−1 compared to the experimental 621.2 mM h−1 with a deviation of 12.94% (Fig. 1D and E). Despite increased deviation, substituents significantly affect electron cloud distribution. Combining the newly tested three catalysts with the initial five and re-conducting the linear surrogate model, the fitting results yielded an R2 value of 0.9828 for eqn (2) (Fig. 1F), confirming the robustness of this catalyst design and predictive modeling strategy. This technique promotes the fine-tuning of catalytic activity, provides guidance for experimental work, lowers research costs, and predicts future research directions. This DFT-based surrogate model is referred to as the “Prediction model of anthraquinone photocatalytic H2O2 production based on electrostatic potential and dipole moment (PMAQ-PHP-EPDM)”.


image file: d4gc02385d-f1.tif
Fig. 1 (A) The structures of five anthraquinone molecules used to establish linear relationships and the corresponding reaction conditions. (B) The dipole moments and generation rates of H2O2 table of basic 5 AQs. (C) Fitting curves and linear regression equation (eqn (1)) derived from the dipole moments of 5 AQs and their correlation with photocatalytic H2O2 production rates. (D) Three anthraquinone catalysts obtained via “Prediction model of anthraquinone photocatalytic H2O2 production based on electrostatic potential and dipole moment” (PMAQ-PHP-EPDM). (E) The dipole moments and generation rates of H2O2 table of total 8 AQs. (F) Fitting curves and linear regression equation (eqn (2)) derived from the dipole moments of 8 AQs and their correlation with photocatalytic H2O2 production rates.

In addition to quantitative predictions, a qualitative assessment was conducted on a wider range of anthraquinone derivatives with diverse functional groups. Electrostatic potential diagrams revealed crucial features contributing to photocatalytic H2O2 production. High-performing compounds like DBAQ, MAQ, BAQ, EAQ, and TAAQ showed a richly electronegative rhombic ring structure characterized by the C[double bond, length as m-dash]O group and adjacent benzene rings (Fig. 2, red diamonds). Notably, these structures lack an electron-deficient center illustrated in deep blue.


image file: d4gc02385d-f2.tif
Fig. 2 Qualitative relationship between the catalysts’ electrostatic potential energy surface and photocatalytic H2O2 production (all data are experimentally measured).

Anthraquinones exhibiting electron-deficient centers (Alizarin, HAQ, Emodin, DAAQ, Fig. 2, blue circles) showed poor performance. This reduced performance is attributable to the increased electronegativity from hydroxyl or amino substitutions, which increased the dipole moments and hindered excited-state hydrogen atom transfer (HAT) processes via hydrogen bonds.

In summary, we have established a strategy for the rapid design, screening, and performance prediction of anthraquinone photocatalysts. First, a series of anthraquinone compounds were drawn using Gaussian View software. Then, electrostatic potential diagrams were obtained based on DFT calculations, and anthraquinone compounds with more red or yellow regions and no blue regions were selected (indicating that AQs with strong electron-rich properties and no electron-deficient centers have better HAT performance). On this basis, molecular dipole moments were calculated via DFT, and the calculated dipole moments were substituted into the linear equation of the existing predictive model (PMAQ-PHP-EPDM, eqn (2)), thereby obtaining the predicted photocatalytic H2O2 production values of these anthraquinone compounds under standard conditions. Catalysts without blue electron-deficient centers have smaller dipole moments and lower molecular polarity, which is conducive to uniform electron distribution and results in better HAT performance.53

The novel anthraquinone catalyst, 2,6-di-tert-butylanthraquinone (DBAQ), features tert-butyl substitutions at the 2,6-positions, enhancing electron cloud density on the ring compared to BAQ. This modification significantly improves HAT and electron transfer (ET) efficiencies under light.50,51 DBAQ synthesis is simple and doesn't require column chromatography purification. Initially, anthracene ring alkylation adds two tert-butyl groups, followed by oxidation to produce DBAQ (see ESI).

After selecting DBAQ as the optimal photocatalyst, we initially investigated the light source's impact on the photocatalytic reaction (Scheme 2A). Two light sources, a 300 W Xe lamp and a 40 W LED, were compared. Despite the Xe lamp's higher power, the H2O2 production rate under standard conditions with simulated sunlight (AM 1.5G, λ > 320 nm) was only 356.8 mM h−1, the low H2O2 production rate can be attributed to the significant decomposition of H2O2 caused by the ultraviolet portion of the light. In contrast, using a 427 nm blue LED light yielded a higher production rate. Subsequently, we explored LED light sources of different wavelengths, focusing on energy efficiency and catalytic efficiency. The wavelength of the light source has been screened, and the optimal wavelength is 427 nm, with a H2O2 production rate of 621.2 mM h−1. Consequently, a 40 W 427 nm Kessil LED lamp was selected as the optimal light source.


image file: d4gc02385d-s2.tif
Scheme 2 Screening and investigation of reaction conditions.

The optimal catalyst loading was determined by reducing DBAQ below the existing 0.1 mol% loading (Scheme 2B). However, this resulted in a noticeable drop in the H2O2 production rate per unit time. Increasing to 0.2 mol% did not lead to a significant improvement in the production rate (592.8 mM h−1). Higher loadings (>0.5 mol%) caused turbidity and a sharp reduction in photocatalytic efficiency. Consequently, 0.1 mol% (4.16 g L−1) was selected as optimal, achieving a hydrogenation efficiency (HE) of 21.0 g L−1 for intermittent H2O2 production. In comparison, the industrial EAQ method uses 140 g L−1 for an HE of 12.0 g L−1. A 75% HE increase was achieved by our system while reducing catalyst usage by 97% (wt%).

To investigate pH's influence, phosphoric acid was used as an acidic modifier. Addition notably enhanced H2O2 production. The highest rate, 699.7 mM h−1, was at 0.1 mol% phosphoric acid, H2O2 mass fraction 3.03 wt%. Further addition decreased H2O2 production. Above 0.5 mol%, rates returned to baseline. In alkaline conditions, 1 mol% KOH yielded 446.8 mM h−1. However, strong alkalinity causes H2O2 instability and decomposition to H2O and O2 (Scheme 2C and D).

Expansion of hydrogen atom donors

The traditional AQ method for H2O2 production involves using hydrogen gas as a hydrogen source and Pd/C as a hydrogenation catalyst under pressure in a fixed bed. Our approach is to activate the catalyst's potential through illumination at very low loadings without requiring H2. This enables efficient hydrogen atom transfer (HAT) and avoids expensive catalysts and reaction risks.

Under intermittent conditions, we explored a total of 16 hydrogen atom donors (HADs) (Scheme 3), including low-carbon alcohols (C < 5) such as ethanol, isopropanol, n-butanol, 1,3-propanediol, and 1,3-butanediol. Apart from isopropanol, alcohols like ethanol (487.1 mM h−1), n-butanol (361.9 mM h−1), and 2-butanol (307.2 mM h−1) exhibited high H2O2 production rates. However, for high-viscosity compounds like diols or aromatic alcohols, magnetic stirring was unable to fully unleash their potential as HADs, resulting in relatively lower H2O2 production rates. Therefore, the development of a reactor that can maintain good lighting conditions while achieving efficient mass transfer of gas–liquid two-phase is crucial for the efficient photocatalytic production of H2O2. This challenge is also a common issue faced by various photocatalytic reactions.


image file: d4gc02385d-s3.tif
Scheme 3 Expansion of hydrogen atom donors. [thin space (1/6-em)]Unless otherwise specified, HADs 1 (3 mL), and DBAQ (0.1 mol%), were added to a quartz beaker equipped with a stirring bar. The mixture was sonicated to dissolve completely. Then, the mixture was stirred under 1.0 atm O2 atmosphere with exposure to 427 nm blue LED at room temperature for 1 h.

Continuous photocatalytic production of H2O2

The photocatalytic amplification effect poses a significant challenge for replacing the industrial anthraquinone process in H2O2 production. As the reaction vessel size increases, light illumination efficiency decreases sharply, affecting gas–liquid mass transfer. Continuous flow strategies in photocatalysis, utilizing microchannel reactors, have gained traction for their efficiency in reducing light exposure distance and liquid layer thickness. Transparent microchannel reactors, like plate, fixed-bed, and coiled-tube types, enhance gas–liquid mass transfer through internal design or fillers, improving specific surface area and generating turbulence. A novel single-pass high-throughput continuous photocatalytic microchannel reactor, using an optimized catalyst, enables efficient continuous H2O2 production.

Establishment of a single-pass high-throughput continuous photocatalytic setup (SPHT-CP setup)

The SPHT-CP setup comprises a high-pressure plunger pump, gas mass flow meter, Y-shaped gas–liquid mixer, gas–liquid pre-cooling module, bead-filled PFA coiled tube reaction module, tail-end cooling module, and air-cooling module. The PFA tube used has an inner diameter of 2.0 mm and an outer diameter of 3.0 mm, with glass beads of 1.5 mm outer diameter used for filling.

Traditional PFA coiled tube reactors achieve gas–liquid mixing at the Y-shaped mixer by adjusting flow rates, followed by plug flow in the coiled tube module. However, gas–liquid mass transfer is limited to surface diffusion segments. To enhance mass transfer, transparent glass beads were added to promote turbulent flow, ensuring continuous gas–liquid mixing and maximizing mass transfer efficiency.

However, for a simple flow channel, filling it with glass beads is relatively straightforward. In our SPHT-CP setup, the total length of the tubing reaches 112 meters (Scheme 4C). Filling the entire length with glass beads in such an extensive tubing system is a challenging task. Nevertheless, we accomplished this by modular filling in four cylindrical continuous micro-packed reactors, each with an outer length of 50 cm and a diameter of 6 cm. These reactors allow for the flexible selection of the reaction channel's length based on the required reaction residence time, ranging from 28 m to 112 m.


image file: d4gc02385d-s4.tif
Scheme 4 The schematic diagrams and physical images of SPHT-CP setup.

The reserved volume of each reactor ranges from 40 mL to 160 mL.

In the realm of continuous microchannel photocatalysis, balancing enhanced gas–liquid mass transfer efficiency with sufficient light intensity for efficient single-pass H2O2 production is challenging. Our system addresses this by enhancing Taylor gas–liquid mass transfer using a PFA reaction tubing filled with glass beads, coiled around a quartz cylinder. Two layers of reflective films inside and outside the reactor, coupled with two 425 nm Kessil light sources, ensure optimal light intensity (64.3 mW cm−2) on the PFA surface for efficient illumination and reaction (Scheme 4B and D).

In the single-pass continuous reactor described above, we conducted an extended study on various HADs that were previously tested under batch conditions. It was observed that, regardless of whether alcohols or toluene and its homologs were used as HADs (Table 1, entries 9 and 10), the production rate of H2O2 per unit time significantly increased in continuous conditions, ranging from 4 to 403-fold improvement. In particular, when isopropanol was used as the HAD, the concentration of H2O2 at the outlet reached 2.01 wt% single-run time of 7 minutes, with a space–time yield of 7.90 [mol (L h)−1]. This corresponds to the production of 107.5 g of a 30 wt% H2O2 solution per hour (pure H2O2 yield = 32.3 g h−1) (Table 1, entry 2). The H2O2 concentration at the outlet already met the requirements of medical disinfectants, surpassing our previous reported photocatalytic systems and currently represents the highest level reported in the field of photocatalytic H2O2 production.

Table 1 Expansion of hydrogen atom donors in continuous flow strategy

image file: d4gc02385d-u1.tif

Entrya 1 Residence time (min) H2O2 (wt%) Generation rate of H2O2 STY
(mM min−1) (mM h−1) [mol (L h)−1]
a Unless otherwise specified, HADs 1 (100 mL), and DBAQ (0.1 mol%), were added to a beaker and sonicated to dissolve completely. Then, the mixture was thoroughly mixed with oxygen through a Y-shaped mixer and then pumped into the photocatalytic reactor at room temperature, vgas[thin space (1/6-em)]:[thin space (1/6-em)]vliq = 28[thin space (1/6-em)]:[thin space (1/6-em)]4 (mL min−1).
1 image file: d4gc02385d-u2.tif 8.5 1.39 38.0 2278.7 4.56
2 image file: d4gc02385d-u3.tif 7.0 2.01 65.8 3950.6 7.90
3 image file: d4gc02385d-u4.tif 12.0 1.04 20.8 1245.9 2.49
4 image file: d4gc02385d-u5.tif 12.0 1.09 21.6 1294.6 2.59
5 image file: d4gc02385d-u6.tif 12.0 2.14 46.2 2772.9 5.55
6 image file: d4gc02385d-u7.tif 14.0 2.71 50.4 3022.1 6.04
7 image file: d4gc02385d-u8.tif 18.0 1.60 27.2 1630.6 3.26
8 image file: d4gc02385d-u9.tif 19.0 0.95 14.9 892.5 1.79
9 image file: d4gc02385d-u10.tif 8.5 0.45 13.4 804.9 1.61
10 image file: d4gc02385d-u11.tif 8.5 0.53 15.9 953.5 1.91


The longer reaction channel caused higher pressure (up to 0.8 MPa) due to the viscous nature of cyclohexanol in continuous conditions. However, by optimizing the reaction liquid composition, efficient hydrogen supply from cyclohexanol-type hydrogen atom donors (HADs) was achieved. The resulting product, KA oil (mixture of cyclohexanol and cyclohexanone), has industrial applications and can be oxidatively ring-opened to obtain adipic acid, crucial for nylon 6 production. In intermittent photocatalytic reactions, the H2O2 yield from cyclohexanol was only 8.54 mM h−1 (Scheme 3, 1g). In continuous flow strategy, by optimizing the cyclohexane[thin space (1/6-em)]:[thin space (1/6-em)]cyclohexanol molar ratio to 2[thin space (1/6-em)]:[thin space (1/6-em)]1 led to a significant increase in H2O2 production rate, reaching 3022.1 mM h−1, which is 403-fold improvement over batch conditions (Table 1, entry 5). This equates to producing 82.2 g of a 30 wt% H2O2 solution per hour (pure H2O2 yield = 24.7 g h−1). H2O2 can be separated using a continuous extraction apparatus,37 and the organic phase containing the catalyst can be recycled, enabling continuous and on-demand H2O2 production.

Moreover, various low-carbon alcohols achieved high H2O2 production efficiency (from 1245.9 to 2278.7 mM h−1) in this continuous flow strategy (Table 1). Higher viscosity HADs also saw significant production rate increases (from 892.5 to 1630.6 mM h−1), showcasing improved gas–liquid mass transfer and photoefficiency in the continuous flow device, leading to enhanced catalyst turnover numbers (TON) and efficient hydrogen atom transfer (HAT) processes for different HADs.

In a 48-hour uninterrupted reaction, a continuous photocatalytic device for H2O2 production was evaluated. Isopropanol served as the hydrogen atom donor, oxygen as the oxygen source, and DBAQ as the photocatalyst at 0.1 mol% loading. The gas–liquid flow rate ratio was 28[thin space (1/6-em)]:[thin space (1/6-em)]2 (ml min−1), and the reaction pressure was 0.3 MPa. For the initial 6 hours, H2O2 concentration in the outlet solution was monitored hourly, with a consistent production rate of 44.2 mM min−1 (2652.0 mM h−1) and sample variance of 0.34 (mM min−1)2 (Fig. 3A). The stable operation allowed sampling every 2 hours from 6 to 48 hours. Results showed stable and efficient H2O2 production throughout, with a production rate of 44.8 mM min−1 (2688 mM h−1) over the entire 48 hours. Sample variance for the production rate from 27 data points was 0.92 (mM min−1)2. The H2O2 mass concentration remained at 1.56 wt%, with a low sample variance of 0.0011.


image file: d4gc02385d-f3.tif
Fig. 3 (A) Concentration and production rate of H2O2 in DBAQ system for continuous photocatalytic preparation of H2O2 for 48 hours in SPHT-CP setup. (B) Impact of reduced reactor pressure on photocatalytic hydrogen peroxide production. (C) Comparison of 1H-NMR spectra of DBAQ before and after photocatalysis. (D) The performance of the recovered DBAQ catalyst in the photocatalytic production of H2O2.

Reducing the pressure within the reactor decreases the compression of oxygen, thereby enhancing the turbulence of the gas–liquid two-phase flow in the reaction pathway. Therefore, at pressures of 0.2 MPa and 0.1 MPa, the H2O2 mass fractions at the outlet increased to 1.82 wt% and 2.24 wt%, respectively. Correspondingly, the H2O2 production rates were elevated to 52.2 mM min−1 and 64.2 mM min−1 (3132.4 mM h−1 and 3852.3 mM h−1) (Fig. 3B).

Recovery of HAD isopropanol and the DBAQ catalyst through rotary evaporation and crystallization. 1H-NMR spectra showed that the catalyst's structure remained unchanged, achieving a recovery rate of 93.6% (Fig. 3C). A continuous photocatalytic experiment was conducted for 4 hours using the recovered DBAQ. The results revealed that the average production of H2O2 within the 4-hour period was 45.2 mM min−1 (Fig. 3A and D). Remarkably, under identical experimental conditions, the H2O2 production remained consistent with the results obtained from the 48-hour continuous experiment. Ultimately, a 310 g working solution of H2O2 with a concentration of 21 wt% was obtained.

Mechanism exploration

Based on previous research, we initially conducted a blank control experiment under standard experimental conditions. The results indicated that, in the absence of light or a catalyst, H2O2 production was almost negligible (Fig. 4A). To further investigate the mechanism of H2O2 generation, we introduced various scavengers into the standard reaction conditions, including the free radical inhibitor TEMPO (2,2,6,6-tetramethylpiperidoxyl), singlet oxygen quencher DABCO (1,4-diazabicyclo[2.2.2]octane), electron scavenger AgNO3, hole scavenger NaC2O4. Experimental results revealed that in the group with the addition of the singlet oxygen quencher, H2O2 production was significantly suppressed, with a H2O2 yield of only 0.5 mM h−1, compared to the control group's 49.1 mM h−1 (Fig. 4B). In the group with the addition of the free radical scavenger TEMPO, the reaction was also partially inhibited, suggesting the generation of free radical intermediates in the H2O2 production process.
image file: d4gc02385d-f4.tif
Fig. 4 (A) Blank control experiment (under intermittent standard conditions). (B) The effect of various scavengers on the photocatalysis process (switching to cyclohexane as the HAD on the basis of intermittent standard conditions). (C) EPR test of the photocatalytic system to detect the generation of singlet oxygen (1O2) under the irradiation. (D) The quench of DBAQ* with different concentrations of isopropanol. (E) Plausible mechanism for the production of H2O2 under DBAQ photocatalytic system.

To further investigate the reaction mechanism, we employed EPR analysis to capture reactive oxygen species (ROS) during the reaction process52 (Fig. 4C). EPR signals corresponding to singlet oxygen radicals was captured using TEMP (2,2,6,6-tetramethylpiperidine). The fluorescence quenching experiment confirmed that as the concentration of isopropanol in the reaction system increased, the fluorescence intensity of the reaction system gradually decreased, indicating that isopropanol can effectively quench the excited state of DBAQ (Fig. 4D). Additionally, high-resolution mass spectrometry was employed to capture critical isopropanol radical intermediates during the course of the reaction (Fig. S4).

In addition, when utilizing EAQ as photocatalyst, TEMPO can capture the respective radical intermediates. However, this method failed to capture the relevant intermediates in the DBAQ photocatalytic H2O2 production system due to DBAQ's superior hydrogen atom abstraction and HAT efficiency in its excited state, preventing chemical capture of the semi-hydrogenated state of the catalyst radical intermediate.37,53

In summary, we propose a possible photocatalytic mechanism for H2O2 generation. In this mechanism, the ground-state photocatalyst DBAQ is excited to its excited state photocatalyst DBAQ* upon exposure to light.36,54 This excited-state photocatalyst can first undergo energy transfer (ET) for quenching and catalyst regeneration. The singlet-state photocatalyst facilitates the activation of triplet oxygen (3O2) during intersystem crossing (ISC), resulting in the more oxidizing singlet oxygen (1O2).

Simultaneously, a portion of the excited-state photocatalyst, through a hydrogen atom transfer process55–57 (HAT), abstracts a hydrogen atom from the hydrogen atom donor, forming the corresponding hydrogenated anthraquinone free radical intermediate (I, HDBAQ). This intermediate rapidly reacts with the photochemically generated singlet oxygen to efficiently produce H2O2. HDBAQ then oxidatively quenches to return to the ground state, completing the catalyst cycle. The HAD, isopropanol, generates a corresponding radical (II) after losing one hydrogen atom, and this radical, upon interaction with one molecule of O2 or 1O2, forms a superoxide free radical intermediate (III). This intermediate, in turn, abstracts an active hydrogen from HDBAQ,1,54 eventually yielding superoxide (IV) and, ultimately, releasing one molecule of H2O2 to form acetone.

Experimental

Synthesis of catalyst

Alkylation of anthracene for the synthesis of 2,6-di-tert-butylanthracene. The synthesis of 2,6-di-tert-butylanthracene38 involved the reflux of anthracene with tert-butyl alcohol in trifluoroacetic acid as a solvent for 24 hours, as depicted in Fig. S1.
Oxidative synthesis of 2,6-di-tert-butylanthraquinone (DBAQ) from 2,6-di-tert-butylanthracene. In a 500 mL round-bottom flask,39 55.0 mmol (15.96 g) of 2,6-di-tert-butylanthraquinone and 160 g of glacial acetic acid were combined. The reaction mixture was then heated to 124 °C and maintained at this temperature for 15 minutes before cooling it to room temperature, as depicted in Fig. S2.

Experimental steps for photocatalytic production of H2O2

Intermittent standard conditions. Precisely measure and transfer the hydrogen atom donor and the catalyst, DBAQ, into a 10 mL vial equipped with a magnetic stirrer. Before initiating the reaction, evacuate the sealed vial to establish a vacuum inside. Connect a syringe filled with pure oxygen to the vial to ensure a pure oxygen environment within the system. Throughout the reaction, employ forced air cooling to maintain room temperature conditions during the reaction. Keep a constant distance of 4.0 cm between the light source and the vial. Activate the light source and the magnetic stirrer to commence the reaction. This photocatalytic H2O2 production strategy is a homogeneous photocatalysis process.

Continuous standard conditions

Single-pass high-throughput continuous photocatalytic setup (SPHT-CP setup). To prepare the photocatalytic reaction working solution, dissolve 1.30 mmol (0.4164 g) of DBAQ in 100 mL of isopropanol using ultrasonication until complete dissolution (0.013 M). The single-pass high-throughput continuous photocatalytic setup (SPHT-CP setup) consists of high-pressure plunger pump, gas mass flow meter, Y-shaped gas–liquid mixer, gas–liquid pre-cooling module, packed-bed PFA coiled tube reaction module, trailing cooling module, an air cooling module and LED light sources. The inner diameter of the PFA tubing used is 2.0 mm, and the outer diameter is 3.0 mm. Glass beads with an outer diameter of 1.5 mm are filled inside the PFA tubing. The PFA tubing is coiled around the outer wall of a quartz tube with a length of 50 cm and an outer diameter of 8 cm. During the reaction, the flow rate of the gas–liquid mixture and reaction time are controlled using a high-pressure plunger pump and a gas mass flow meter.

Titration analysis of H2O2

Titration reaction procedure36,37. Take a small sample of the reaction mixture in a beaker and accurately measure its mass. Add 8 ml of 2 M sulfuric acid solution, followed by 5 drops of ammonium molybdate solution to enhance color development. Next, add 3 ml of starch indicator and 6 ml of potassium iodide solution. If H2O2 is produced in the reaction, the solution will turn dark purple. Finally, titrate the test solution with a standard sodium thiosulfate solution, shaking the beaker while adding the titrant until the solution fades when adding the last half-drop. Record the volume of the standard solution used and calculate the concentration of H2O2. The specific titration reaction equation and the concentration of H2O2 are shown in eqn (a)–(d).40,41
 
2H+ + H2O2 + 2I ↔ 2H2O + I2(a)
 
2S2O32− + I2 ↔ S4O62− + 2I(b)
 
image file: d4gc02385d-t1.tif(c)
 
image file: d4gc02385d-t2.tif(d)

Conclusion

The PMAQ-PHP-EPDM model is developed for AQ catalyst designs based on DFT calculations and validated through the synthesis and application of DBAQ. DBAQ achieved the highest reported H2O2 production rates in both batch (621.2 mM h−1) and continuous (3950.6 mM h−1) conditions at a low catalyst loading of 0.1 mol%. In our self-constructed high-throughput single-pass continuous reactor (SPHT-CP setup), using a working solution comprising cyclohexanol and cyclohexane (molar ratio 2[thin space (1/6-em)]:[thin space (1/6-em)]1) with DBAQ, at a retention time of 14.0 minutes, the outlet H2O2 concentration reached 2.71 wt%, with a space–time yield of 6.04 (mol (L h)−1) and a turnover number (TON) of 112. This efficient continuous photocatalytic H2O2 production strategy improved stability and catalytic efficiency while achieving single-pass continuous H2O2 production, offering a solution for on-demand synthesis. It also introduced working solutions for oxidation platforms and a modular continuous oxidation setup, advancing research on higher-value-added products.

Data availability

The authors confirm that the data supporting the findings of this study are available within the article [and/or] its ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We are grateful for financial support from the National Natural Science Foundation of China (U20A20143). We would like to thank the State Key Laboratory of Fine Chemicals and the Fundamental Research Funds for China Central Universities (DUT22LAB608).

References

  1. J. M. Campos-Martin, G. Blanco-Brieva and J. L. G. Fierro, Angew. Chem., Int. Ed., 2006, 45, 6962–6984 CrossRef PubMed .
  2. X. Zeng, Y. Liu, X. Hu and X. Zhang, Green Chem., 2021, 23, 1466–1494 RSC .
  3. R. J. Lewis, K. Ueura, X. Liu, Y. Fukuta, T. Qin, T. E. Davies, D. J. Morgan, A. Stenner, J. Singleton, J. K. Edwards, S. J. Freakley, C. J. Kiely, L. Chen, Y. Yamamoto and G. J. Hutchings, ACS Catal., 2023, 13, 1934–1945 CrossRef .
  4. P. Tang, B. Ji and G. Sun, Carbohydr. Polym., 2016, 147, 139–145 CrossRef PubMed .
  5. Q. Chen, Chem. Eng. Process., 2008, 47, 787–792 CrossRef .
  6. H. Li, B. Zheng, Z. Pan, B. Zong and M. Qiao, Front. Chem. Sci. Eng., 2018, 12, 124–131 CrossRef .
  7. T. Freese, J. T. Meijer, B. L. Feringa and S. B. Beil, Nat. Catal., 2023, 6, 553–558 CrossRef .
  8. H. Tan, P. Zhou, M. Liu, Q. Zhang, F. Liu, H. Guo, Y. Zhou, Y. Chen, L. Zeng, L. Gu, Z. Zheng, M. Tong and S. Guo, Nat. Synth., 2023, 2, 557–563 CrossRef .
  9. X. Zhang, H. Su, P. Cui, Y. Cao, Z. Teng, Q. Zhang, Y. Wang, Y. Feng, R. Feng, J. Hou, X. Zhou, P. Ma, H. Hu, K. Wang, C. Wang, L. Gan, Y. Zhao, Q. Liu, T. Zhang and K. Zheng, Nat. Commun., 2023, 14, 7115 CrossRef CAS .
  10. M. Volokh and M. Shalom, Nat. Catal., 2021, 4, 350–351 CrossRef CAS .
  11. Z. Teng, Q. Zhang, H. Yang, K. Kato, W. Yang, Y. Lu, S. Liu, C. Wang, A. Yamakata, C. Su, B. Liu and T. Ohno, Nat. Catal., 2021, 4, 374–384 CrossRef CAS .
  12. X. Wang, P. Zhou, Q. Zhou, Q. Zhang, H. Ning, M. Wu and W. Wu, Chem. Eng. J., 2023, 476, 146488 CrossRef CAS .
  13. Z. Wang, Y. Zhao, Y. Zhou, X. Wang, H. Huang, Y. Liu, M. Shao and Z. Kang, Chem. Eng. J., 2022, 427, 131972 CrossRef CAS .
  14. Y. Shiraishi, S. Kanazawa, D. Tsukamoto, A. Shiro, Y. Sugano and T. Hirai, ACS Catal., 2013, 3, 2222–2227 CrossRef CAS .
  15. L. Wang, J. Zhang, Y. Zhang, H. Yu, Y. Qu and J. Yu, Small, 2022, 18, 2104561 CrossRef CAS PubMed .
  16. T. Liu, Z. Pan, J. J. M. Vequizo, K. Kato, B. Wu, A. Yamakata, K. Katayama, B. Chen, C. Chu and K. Domen, Nat. Commun., 2022, 13, 1034 CrossRef CAS PubMed .
  17. H. Hou, X. Zeng and X. Zhang, Angew. Chem., Int. Ed., 2020, 59, 17356–17376 CrossRef CAS PubMed .
  18. R. Wang, M. S. Shi, F. Y. Xu, Y. Qiu, P. Zhang, K. L. Shen, Q. Zhao, J. G. Yu and Y. F. Zhang, Nat. Commun., 2020, 11, 4465 CrossRef CAS PubMed .
  19. R. Ma, L. Wang, H. Wang, Z. Liu, M. Xing, L. Zhu, X. Meng and F. Xiao, Appl. Catal., B, 2019, 244, 594–603 CrossRef CAS .
  20. Y. Isaka, Y. Kondo, Y. Kawase, Y. Kuwahara, K. Mori and H. Yamashita, Chem. Commun., 2018, 54, 9270–9273 RSC .
  21. S. Wang, Z. Xie, D. Zhu, M. Bonn, H. I. Wang, Q. Liao, H. Xu, X. Chen and C. Gu, Nat. Commun., 2023, 14, 6891 CrossRef CAS .
  22. Y. Shiraishi, T. Takii, T. Hagi, S. Mori, Y. Kofuji, Y. Kitagawa, S. Tanaka, S. Ichikawa and T. Hirai, Nat. Mater., 2019, 18, 985–993 CrossRef CAS PubMed .
  23. Y. Su, K. Urayama, S. Kitagawa, L. Huang, S. Furukawa and C. Gu, J. Am. Chem. Soc., 2024, 146, 15479–15487 CrossRef CAS .
  24. X. Xu, Y. Sui, W. Chen, W. Huang, X. Li, Y. Li, D. Liu, S. Gao, W. Wu, C. Pan, H. Zhong, H. Wen and M. Wen, Appl. Catal., B, 2024, 341, 123271 CrossRef CAS .
  25. C. Chu, D. Huang, Q. Zhu, E. Stavitski, J. A. Spies, Z. Pan, J. Mao, H. L. Xin, C. A. Schmuttenmaer, S. Hu and J. Kim, ACS Catal., 2019, 9, 626–631 CrossRef CAS .
  26. Q. Zhi, W. Liu, R. Jiang, X. Zhan, Y. Jin, X. Chen, X. Yang, K. Wang, W. Cao, D. Qi and J. Jiang, J. Am. Chem. Soc., 2022, 144, 21328–21336 CrossRef CAS .
  27. M. Deng, J. Sun, A. Laemont, C. Liu, L. Wang, L. Bourda, J. Chakraborty, K. Van Hecke, R. Morent, N. De Geyter, K. Leus, H. Chen and P. Van Der Voort, Green Chem., 2023, 25, 3069–3076 RSC .
  28. F. Hao, C. Yang, X. Lv, F. Chen, S. Wang, G. Zheng and Q. Han, Angew. Chem., Int. Ed., 2023, 62, e202315456 CrossRef CAS .
  29. Q. Liao, Q. Sun, H. Xu, Y. Wang, Y. Xu, Z. Li, J. Hu, D. Wang, H. Li and K. Xi, Angew. Chem., Int. Ed., 2023, 62, e202310556 CrossRef CAS .
  30. Y. Isaka, Y. Kawase, Y. Kuwahara, K. Mori and H. Yamashita, Angew. Chem., Int. Ed., 2019, 58, 5402–5406 CrossRef CAS .
  31. Y. Zhang, C. Pan, G. Bian, J. Xu, Y. Dong, Y. Zhang, Y. Lou, W. Liu and Y. Zhu, Nat. Energy, 2023, 8, 361–371 CrossRef CAS .
  32. A. Gopakumar, P. Ren, J. Chen, B. V. M. Rodrigues, H. Y. V. Ching, A. Jaworski, S. V. Doorslaer, A. Rokicińska, P. Kuśtrowski, G. Barcaro, S. Monti, A. Slabon and S. Das, J. Am. Chem. Soc., 2022, 144, 2603–2613 CrossRef CAS PubMed .
  33. L. Xu, Y. Liu, L. Li, Z. Hu and J. C. Yu, ACS Catal., 2021, 11, 14480–14488 CrossRef CAS .
  34. M. Ren, M. Mao, X. Duan and Q. Song, J. Photochem. Photobiol., A, 2011, 217, 164–168 CrossRef CAS .
  35. G. Xu, Y. Liang and F. Chen, J. Mol. Catal. A: Chem., 2016, 420, 66–72 CrossRef CAS .
  36. H. B. Vibbert, C. Bendel, J. R. Norton and A. J. Moment, ACS Sustainable Chem. Eng., 2022, 10, 11106–11116 CrossRef CAS .
  37. Z. Yu, J. Zhao, Y. Wu, L. Guo, C. Ma, H. Zhu, J. Li, L. Duan, Z. Liu, H. Sun, G. Zhao and Q. Meng, ACS Sustainable Chem. Eng., 2023, 11, 13320–13332 CrossRef CAS .
  38. J. Lee, Y. Chen, J. S. Lin, C. Wu, C. Chen, C. Dai, C. Chao, H. Chen and W. Liau, Tetrahedron, 2011, 67, 1696–1702 CrossRef CAS .
  39. M. V. N. Raju, M. E. Mohanty, P. R. Bangal and J. R. Vaidya, J. Phys. Chem. C, 2015, 119, 8563–8575 CrossRef .
  40. Y. Wei, J. Zhang, Q. Zheng, J. Miao, P. J. Alvarez and M. Long, Chemosphere, 2021, 279, 130556 CrossRef CAS .
  41. N. V. Klassen, D. Marchington and H. C. E. McGowan, Anal. Chem., 1994, 66, 2921–2925 CrossRef CAS .
  42. J. S. Murray and P. Politzer, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2011, 1, 153–163 CAS .
  43. N. Argaman and G. Makov, Am. J. Phys., 2000, 68, 69–79 CrossRef CAS .
  44. P. Politzer and J. S. Murray, Theor. Chem. Acc., 2002, 108, 134–142 Search PubMed .
  45. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson and H. Nakatsuji, Gaussian Inc., Wallingford CT, 2016, 1, 572 .
  46. T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed .
  47. A. St Amant, W. D. Cornell, P. A. Kollman and T. A. Halgren, J. Comput. Chem., 1995, 16, 1483–1506 CrossRef CAS .
  48. J. Zhang and T. Lu, Phys. Chem. Chem. Phys., 2021, 23, 20323–20328 RSC .
  49. T. Lu and Q. Chen, J. Mol. Model., 2020, 26, 315 CrossRef CAS .
  50. J. E. Bachman, L. A. Curtiss and R. S. Assary, J. Phys. Chem. A, 2014, 118, 8852–8860 CrossRef CAS PubMed .
  51. L. Capaldo and D. Ravelli, Eur. J. Org. Chem., 2017, 2056–2071 CrossRef CAS .
  52. Y. Nosaka and A. Y. Nosaka, Chem. Rev., 2017, 117, 11302–11336 CrossRef CAS .
  53. Organic Photochemistry, ed. A. Padwa, CRC Press, 1st edn, 1991,  DOI:10.1201/9780203744826 .
  54. D. Jiang, Q. Zhang, L. Yang, Y. Deng, B. Yang, Y. Liu, C. Zhang and Z. Fu, Renewable Energy, 2021, 174, 928–938 CrossRef CAS .
  55. L. Capaldo and D. Ravelli, Eur. J. Org. Chem., 2017, 2056–2071 CrossRef CAS .
  56. C. Tang, X. Qiu, Z. Cheng and N. Jiao, Chem. Soc. Rev., 2021, 50, 8067–8101 RSC .
  57. L. Capaldo, D. Ravelli and M. Fagnoni, Chem. Rev., 2022, 122, 1875–1924 CrossRef CAS PubMed .

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4gc02385d
Equivalent to first author.

This journal is © The Royal Society of Chemistry 2024