Vishesh
Kumar
a,
Ved
Vyas
a,
Deepak
Kumar
a,
Ashish Kumar
Kushwaha
b and
Arindam
Indra
*a
aDepartment of Chemistry, Indian Institute of Technology (BHU), Varanasi, UP 221005, India. E-mail: arindam.chy@iitbhu.ac.in; Tel: +919919080675
bDepartment of Chemistry, Banaras Hindu University, Varanasi, UP 221005, India
First published on 29th August 2024
Herein, we report the modulation of the band structures of halide perovskite Cs2CuBr4 by tuning the synthesis methods. The photocatalyst PC-1, synthesized by the hot injection method, has a more negative conduction band minima (CBM) than the photocatalyst PC-2, synthesized at room temperature. As a result, PC-1 can activate molecular O2 more efficiently to initiate the radical-mediated dehydrogenation of alcohols. The more positive valence band maxima (VBM) of PC-1 also facilitates amine oxidation to the corresponding radical. Further, improved charge separation and transport and a decrement in the photogenerated charge carrier recombination have been detected for PC-1 to enhance photocatalytic activity. PC-1 showed improved yields for a series of structurally diverse amides (highest yield = 98%) by oxidative amidation of alcohols under visible light irradiation.
In addition, photocatalytic processes using Ru- and Ir-based photocatalysts have been widely explored.2,12–14 Organo-photocatalysts such as phenazine salts,15,16 rose Bengal,17 quinolinium compounds,16 acridinium salts,16etc. have also been reported.16 Although improved activity and selectivity for amides have been achieved with these catalyst systems, the widespread application is often hindered by the requirement of expensive metal and ligand, costly reagents, and poor recycling capacity.
Recently, alternative approaches for the oxidative amidation of aldehydes using visible-light semiconductors have been reported with improved recycling performance. For example, heterogeneous semiconductor catalysts like Ag/g-C3N4, Ni/g-C3N4, TiO2, Mn3O4, Pd/MIL-101(Fe), Fe3O4/PDA/CdS, TzPm-COF, [Ni(dmgH)2]-CsPbBr3, etc. have been explored for oxidative amidation of aldehydes with amines.14,16,18–20 However, these methods often rely on photocatalysts having fewer active sites, fast charge recombination, and limited optical properties.
In contrast, amide bond formation involving the reaction of benzyl alcohol and amine presents an appealing alternative. The efficiency of the dehydrogenative coupling reaction can be improved using alcohols as the substrates instead of aldehydes. Alcohols are largely abundant, less toxic, and less expensive feedstocks, and their direct use makes the process more sustainable.
Milstein group first demonstrated the amidation of alcohols by hydrogen atom transfer (HAT) using ruthenium pincer complex as the catalyst.21 Later, the HAT process was explored for the amidation of the alcohols using transition metal and noble metal-based catalysts.22,23 However, these processes suffer from drawbacks like the requirement of high temperature and the use of excessive oxidants and bases. In this context, Yin group explored photocatalytic approach for the amidation of alcohols using Ag2O/P–C3N4.13
On the other hand, halide perovskites have emerged as promising photocatalysts for different energy conversion processes24–26 and organic photoredox reactions because of their unique optoelectronic properties, such as tunable bandgaps, high charge carrier mobility, and low exciton binding energy.24,27–30 Although Pb-based perovskites, particularly CsPbBr3, have been extensively studied for photocatalytic energy conversion and organic transformation reactions, the high toxicity of lead and the sensitivity of the perovskites to water limits their further applications.31 This motivates the search for lead-free perovskites with high photoredox activity.
Previously, Dong et al. reported superior photocatalytic activity of Cs2CuBr4 compared to CsPbBr3 for photocatalytic CO2 reduction reaction.32 Copper, a 3d-metal with unique physicochemical properties, offers multiple valences to optimize the d-band structure. Further, the adjacent Lewis acid (Cu2+) and Lewis base (Br−) sites in the surface of Cs2CuBr4 can generate surface-frustrated Lewis pair sites-which optimize adsorption of the substrate on the catalyst surface.33–36
These fascinating properties of Cs2CuBr4 have prompted us to explore its photoredox activity for the organic transformation reaction. Further, we have modulated the band structure of Cs2CuBr4 by varying the synthesis methods (Scheme 1). The hot injection method produces PC-1 with a smaller size of nanoparticles compared to room temperature method (PC-2). The effect of the size of Cs2CuBr4 nanoparticles is found to be pronounced for the bandgap (PC-1: 1.89 eV vs. PC-2: 1.64 eV) and band positions (CBMPC-1: −0.80 eV vs. CBMPC-2: −0.72 eV and VBMPC-1: 1.09 eV vs. VBMPC-2: 0.92 eV). Further, PC-1 showed better charge separation and transport of photo-generated charge carriers compared to PC-2. In addition, the more positive VBM of PC-1 (than PC-2) facilitates the formation of cationic radicals from amines by hole transfer.
Scheme 1 Schematic representation showing the structural modulation of Cs2CuBr4 by two different synthesis methods: (i) hot-injection (PC-1) and (ii) room temperature (PC-2). |
The conduction band minima of PC-1 and PC-2 are sufficiently negative (−0.80 V vs. NHE) for the reduction of O2 to O2˙− (−0.33 V vs. NHE). Therefore, the electron transfer from the conduction band of Cs2CuBr4 to the LUMO of molecular O2 favors the formation of superoxide radicals. Previously, the activation of molecular O2 and its interaction with lead halide perovskites has been reported by different groups.37 These studies have revealed that the molecular O2 is adsorbed on the surface of perovskite by weak van der Waals forces and the electron transfer from the perovskite to O2 results in the formation of a strong Pb–O bond.38 This leads to the structural damage of the perovskite. However, an increase in the photoluminescence life-time and emission quantum yield of perovskite in the presence of O2 was observed.39 Therefore, the optoelectronic properties of the perovskite are significantly tuned in the presence of O2 to facilitate the charge-transfer dynamics.
Although both the photocatalysts (PC-1 and PC-2) can activate molecular O2 to superoxide radicals, the optimized bandgap, band positions, and the presence of the surface ligands in PC-1 facilitate the O2 reduction process. As a result, PC-1 achieved high efficiency for the oxidative amidation of alcohols with amines to form structurally diverse amides (highest yield = 98%). Moreover, the catalyst PC-1 can be recycled five times with a minimum loss of initial activity.
The powder X-ray diffraction pattern (PXRD) confirmed the orthorhombic crystal structure of Cs2CuBr4 (JCPDS: 71-1462) with space group Pnma (Fig. S2†).32 The PXRD peaks of PC-1 are positively shifted compared to that of PC-2.
The scanning electron microscopy (SEM) detected large particles (>500 nm) with irregular morphology for PC-2 and rod-shaped particles for PC-1 (Fig. 1a and S3†). Transmission electron microscopy (TEM) also revealed the rod-shape morphology of PC-1 with size ranging from 100–200 nm (Fig. 1b).32,40,41 High-resolution TEM (HRTEM) confirmed the lattice spacing of 0.32 nm, corresponding to the (203) plane of Cs2CuBr4 (Fig. 1c). Further, Fast Fourier transformed (FFT) and inverse FFT revealed the (203) plane of Cs2CuBr4 (Fig. 1d and d′). The EDX spectrum of PC-1 and PC-2 confirmed the presence of Cs, Cu, and Br (Fig. S4 and S6†). The EDX elemental mapping also detected a uniform distribution of the elements Cs, Cu, and Br (Fig. S5 and S7†).
The variation in the synthetic method of Cs2CuBr4 not only shifts the PXRD peaks, it also tunes the electronic structure of the elements as detected by X-ray photoelectron spectroscopy (XPS). The Cs 3d XPS was deconvoluted into two peaks for Cs 3d5/2 and Cs 3d3/2 (Fig. 1e). A positive shift of the Cs 3d5/2 peak by 0.49 eV was detected for PC-1, indicating a significant variation in the electron density around Cs.
Further, the Cu 2p XP spectra were fitted for Cu 2p3/2 and Cu 2p1/2 peaks (Fig. 1f).32,40,41 Similar to Cs, the positive shift of the 2p3/2 peak by 0.62 eV was detected for Cu. The Cu2+/Cu+ ratio of PC-1 and PC-2 was calculated to be 2.65 and 1.31, respectively. PC-1 has a higher Cu2+/Cu+ ratio compared to PC-2. The modulation of the electronic structure was also confirmed from the Cu 2p3/2-2p1/2 spin–orbit coupling values for PC-1 (21.69 eV) and PC-2 (22.24 eV) respectively. The higher Cu2+/Cu+ ratio in PC-1 improved the Lewis acid character to improve the adsorption of the substrates.
Br 3d peak was also shifted to higher binding energy (0.53 eV) for PC-1 than PC-2 (Fig. 1g). The electronic structure modulation leads to a significant variation in the photocatalytic activity for oxidative amidation of alcohols with amines (see later).
The Mott–Schottky plots displayed positive slopes within the frequency range of 0.5–1.5 kHz, indicating the n-type nature of both PC-1 and PC-2 (Fig. 2b and S8†).32,44,45 The Fermi level energy (Ef) of PC-1 and PC-2 was found to be −0.70 and −0.62 eV. The CBM for n-type semiconductors is 0.1 or 0.2 eV higher than the flat band potential (see details in ESI†).46 Therefore, the CBMs for PC-1 and PC-2 were determined to be −0.80 and −0.72 V vs. the normal hydrogen electrode (NHE), respectively (Fig. 2c). The VBMs of PC-1 and PC-2 were calculated to be 1.09 and 0.92 eV vs. NHE, respectively.
The electronic band structure plays a pivotal role in regulating photocatalytic activity. PC-1 exhibited a wider band gap (1.89 eV) than that of PC-2 (1.64 eV). The more negative CBM of PC-1 (than PC-2) also facilitates the activation of molecular oxygen-forming superoxide radicals to initiate the C–N coupling reaction for the amide formation (see later).
Further, cyclic voltammetry (CV) measurements were carried out to understand the electronic structure of the catalysts. As the redox peak positions in the CV are not prominent, we have carried out differential pulse voltammetry (DPV) in the dark and light (Fig. S9†). Interestingly, the peak positions for PC-1 in the dark and light did not show any significant variation (Fig. S9†). However, the highly negative reduction peaks in PC-1 compared to PC-2 helps in the facile activation of molecular oxygen during the photocatalytic process.
Photoluminescence (PL) spectra showed an emission peak at 460 nm for both the catalysts and the intensity of the emission peak was found to be significantly lower for PC-1 (Fig. 2d).47–50 The photocurrent measurements confirmed better charge separation and transport in PC-1 compared to PC-2 (Fig. 2e).51–54 The charge transport properties of the catalysts were further investigated through electrochemical impedance spectroscopy (EIS) studies and in the presence of light, low charge transfer resistance was detected for PC-1 (Fig. 2f).32,40,41,43–45 These results indicate that smaller particle size of PC-1 results in a better charge separation and transport than that in PC-2 to improve the photocatalytic oxygen activation.
a Reaction conditions: alcohol (0.5 mmol), amine (1.0 mmol), photocatalyst (10 mg), THF (3 mL), 15 W blue LED, temperature: 35 ± 2 °C, time: 10 h. In all the cases, isolated yield of the product was reported and 1H and 13C NMR was used to characterize the products. |
---|
The role of molecular oxygen is crucial for the amidation reaction and the formation of amide is not detected in the presence of N2. The molecular O2 (in air) is reduced to superoxide radicals by the transfer of photogenerated electrons from the CB of Cs2CuBr4 (Fig. 2c). Further, superoxide radicals initiate the dehydrogenation of benzyl alcohol. Therefore, in the absence of O2, the amide formation was not observed. The reaction was found to be faster in pure O2 compared to air (Table 1). The yield of 4-nitrophenyl(pyrrolidine-1-yl)methanone was extremely low (4%) in the dark and no reaction took place in the absence of catalyst. Among different solvents, the best result was obtained in tetrahydrofuran (THF) (Table S2†).
a Reaction conditions: alcohol (0.5 mmol), amine (1.0 mmol), PC-1 (10 mg), THF (3 mL), 15 W blue LED, air, temperature: 35 ± 2 °C, time: 10 h. In all the cases, isolated yield of the product was reported and the products were characterized by 1H and 13C NMR spectroscopy (Table S3). |
---|
In contrast, the electron-withdrawing group at the para-position of benzyl alcohol produced slightly higher yield of amides (Table 2, 3e and 3h). Even, 4-chlorobenzylalcohol produced a good yield of amide (95%) under similar reaction conditions (Table 2, 3g). The trifluoromethyl group at para-position of the phenyl ring also produced a high yield of amide (Table 2, 3h).
The reaction of 2-nitrobenzyl alcohol with pyrrolidine produced 3f, a well-known insect repellent, with 91% yield.55 However, the electron-withdrawing group (–NO2 and –CF3) at the ortho-position of benzyl alcohol produced a slightly lower yield of amide compared to that in the para-position, most probably due to the steric crowding at the ortho-position (Table 2, 3f, and 3i).
The electron-donating or withdrawing properties of the substituents at different positions (ortho-, meta-, or para-) of the phenyl ring of benzyl alcohol may have a complicated impact on the conversion efficiency and product yield. For example, the electron-donating group at the para-position of benzyl alcohol can facilitate the Cα–H abstraction by O2˙− but the electron-withdrawing group can stabilize the intermediate benzylic radical for a facile C–N coupling.
The photocatalytic amidation was also found to be effective for the heterocyclic alcohols to produce more than 82% yield of the amides. However, the amide yields are low for heterocyclic alcohols compared to benzyl alcohol (Table 2, 3j and 3k).
The variation of the amine also affected the amide yields. The amines like 2-oxazolidine or morpholine, when reacted with 4-nitro benzyl alcohol, a significant decrease in the amide yield was observed compared to pyrrolidine (Table 2, 3evs.3l and 3n) while piperidine produced similar yield of amide (Table 2, 3m).
Interestingly, heterocyclic amine (2-aminopyridine) yielded 76% amide when reacted with 4-nitrobenzyl alcohol (Table 2, 3o). However, aniline reduced the yield of amide (Table 2, 3p). Further, benzylamine produced 71% yield of amide reacting with benzyl alcohol (Table 2, 3q). However, amide yield was significantly reduced when 4-nitrobenzyl alcohol was reacted with open-chain aliphatic amine (diethyl amine) (Table 2, 3r).
The above results showed that the yields of amides for aliphatic cyclic amines are higher than that for aliphatic acyclic amines, aromatic primary amines, and hetero-aromatic amines. The poor nucleophilic nature of aromatic primary amine and hetero-aromatic amine and the steric crowding in aliphatic acyclic amine reduced the amide yield.7,19 The presence of strong electronegative oxygen in the ring of aliphatic cyclic amine (morpholine and 2-oxazolidine) also reduced the nucleophilic character of it and hence a low product yield was obtained.55 The catalyst stability was evaluated by recycling 5 cycles with a minimum loss of photocatalytic activity (Fig. S10†). After 5 cycles of oxidative amidation of 4-nitrobenzyl alcohol, no significant change in the UV-visible-DRS spectrum of PC-1 was observed (Fig. S11†). The IR spectroscopy also detected the presence of the surface oleylamine in PC-1 after the first catalytic cycle (Fig. S12a†).
The effect of surface ligand on the photocatalytic reaction is complex. The surface ligand reduces the non-radiative recombination process and improves the system's photoluminescence property, and hence can improve the photocatalytic activity. On the other hand, the ligand can also passivate the catalyst surface, inhibiting the substrate binding.56,57 As the surface ligand plays an important role in charge separation and transport of the photocatalyst, we have checked the activity of PC-1 after the complete removal of the ligand oleylamine from the catalyst surface (see in the experimental part, ESI) (Fig. S12b).† The removal of surface ligand from PC-1 leads to a slight decrease in the yield (87%) of the amide.
The process of C–N bond formation is governed by two steps: (i) the single electron reduction of O2 to O2˙− radicals and (ii) oxidation of the amine to the corresponding cationic radicals (2a*) by the holes (h+) in the valence band. As PC-1 has a more negative CBM (ECB = −0.80 V vs. NHE, pH = 7) than PC-2, the single electron reduction of O2 to O2˙− (O2 + e− → O2˙−, Eo = – 0.33 V vs. NHE, pH = 7) is favored for the former.19 Further O2˙− initiates the dehydrogenation of 4-nitrobenzyl alcohol and form benzyl radical (1e*) and ˙OOH. The bond dissociation energy of Cα–H bond (79.3 kcal mol−1) in benzyl alcohol is lower than that of the O–H bond (99.0 kcal mol−1).58 Therefore, the photogenerated O2˙− radicals abstract the hydrogen from Cα–H bond.59 In the next step, the radical coupling of 1e* and 2e* forms an intermediate (I#) species. The ˙OOH radicals further abstract proton from I# to form A* and hydrogen peroxide. Further reaction of O2˙− with A* forms final product 3e and H2O2 (Scheme 2).
Scheme 2 The proposed mechanism for photocatalytic oxidative amidation of alcohol is based on the detected intermediates. |
In contrast to the previous reports, we are not able to detect the formation of 4-nitrobenzaldehyde by the oxidation of 4-nitrobenzyl alcohol in the presence of pyrrolidine. However, in the absence of amine, 4-nitrobenzyl alcohol is slowly converted into the corresponding aldehyde (Fig. S13†). Further, we have performed the reaction of 4-nitrobenzaldehyde with pyrrolidine and the rate of the reaction was found to be slower than that of 4-nitrobenzyl alcohol and pyrrolidine. The reaction rate for the oxidation of 4-nitrobenzyl alcohol (without pyrrolidine) to 4-nitrobenzaldehyde was found to be slower than the C–N bond formation but faster than the reaction of 4-nitrobenzaldehyde with pyrrolidine (Fig. S13†). The above experiments showed that the photocatalytic reaction proceeds through the direct coupling of 4-nitrobenzyl alcohol and pyrrolidine instead of the benzaldehyde route. These studies confirmed the radical coupling of 1e* and 2e* to form I# compared to the oxidation of 1e to 4-nitrobenzaldehyde.
A series of quenching experiments were carried out to understand the involvement of the electrons from the CB, holes from the VB, and O2˙− radicals (Fig. S14†). The addition of AgNO3 and triethylamine as the electron and hole scavenger resulted in a significant drop in the yield of 3e to 12% and 15%, respectively.60 In addition, when 4-benzoquinone (BQ) was used as a superoxide radical (O2˙−) scavenger, the yield of amide decreased to 21% (Fig. S14†).58,61,62
The involvement of the radical mechanism was also proved by radical trapping reaction. In the presence of TEMPO (2,2,6,6-Tetramethylpiperidine-1-oxyl), the yield of 3e was reduced to 18% (Fig. S15†). Further, we are able to detect TEMPO-1e* and TEMPO-2e* adducts by mass spectrometry (Fig. S15†). The production of O2˙− was detected by p-nitro-blue tetrazolium chloride (NBT) test (Fig. S16†).63 The UV-visible spectroscopy confirmed (NBT test) that PC-1 produced more O2˙− compared to PC-2.64
The formation of H2O2 was also detected by using o-tolidine as an indicator (Fig. S17†).65–67 Therefore, it is clear that PC-1 is more effective in the activation of molecular oxygen compared to PC-2 (Fig. S16 and S17†). Overall, the above study revealed that PC-1 nanoparticles showed improved photocatalytic activity compared to that of PC-2.
According to previous reports, copper(II) sites in the catalyst facilitate the electron transfer process and stabilize the reactive intermediates formed during the coupling reaction, thereby influencing the reaction kinetics and selectivity of the products.68 Cu(I) sites typically initiate C–N coupling reactions.69,70 Therefore, the presence of both Cu(II) and Cu(I) sites in PC-1 is beneficial for the selective C–N coupling of alcohol and amine.
Footnote |
† Electronic supplementary information (ESI) available: Catalyst synthesis, characterizations, optimization of the catalytic reactions, and 1H NMR and 13C NMR spectra are available. See DOI: https://doi.org/10.1039/d4sc03796k |
This journal is © The Royal Society of Chemistry 2024 |