Ruchika
Gupta
a,
Aashish
a,
Upma
a,
Shubhangi
Majumdar
b,
Pramit K.
Chowdhury
b and
Rajeev
Gupta
*a
aDepartment of Chemistry, University of Delhi, Delhi 110 007, India. E-mail: rgupta@chemistry.du.ac.in; Web: https://people.du.ac.in/%7Ergupta/
bDepartment of Chemistry, Indian Institute of Technology, Delhi, India
First published on 14th October 2024
The utilization of earth-abundant photosensitizers with visible light absorption to enable sustainable photocatalysis is a long-standing challenge. Overcoming such a challenge, in this work, two lanthanide (Ln3+ = Tb, Eu) based metal–organic frameworks (Ln-MOFs) have been synthesized utilizing a Co3+-based metalloligand. Both Ln-MOFs function as remarkable photocatalysts for the selective oxidation of assorted alcohols and sulfides to their corresponding aldehydes/ketones and sulfoxides using visible light. The photophysical behavior of both Ln-MOFs and mechanism of photocatalysis is comprehensively investigated using time-resolved transient absorption spectroscopy, electrochemical impedance spectroscopy, electron paramagnetic resonance spectroscopy, photoluminescence and phosphorescence studies. In both Ln-MOFs, a metalloligand acts as a light-harvester, being excited by visible light, while Ln3+ ions endow the resulting MOFs with long-lived triplet excited states. Ultrafast transient absorption spectroscopy, further supported by electron paramagnetic resonance spectra, revealed excited-state electron transfer from metalloligands to the Ln3+ ions and transient generation of Ln2+ sites alongside the facilitation of intersystem crossing. The excited Ln2+ ions transfer energy to the ground-state triplet oxygen (3O2) to generate singlet oxygen (1O2). The HOMO–LUMO positions of both Ln-MOFs support the generation of ˙O2− and 1O2 but inhibit strongly-oxidizing yet non-selective ˙OH radicals. Scavenger experiments, 1O2 traps and electron paramagnetic resonance spectra confirmed the generation of singlet oxygen. The heavy-metal effect of a lanthanide ion in Ln-MOFs for the generation of triplet excitons is confirmed by the synthesis of a non-heavy-metal analogue involving a zinc ion via a single-crystal-to-single-crystal transformation strategy. The present results are noteworthy and may aid in the development of other earth-abundant metalloligand-based photocatalysts for challenging yet sustainable catalysis.
Metal–organic frameworks (MOFs) have attracted significant interest as they offer several notable features including a chemically stable and thermally robust network, high crystallinity, high porosity, large surface area and architectural manoeuvrability.29–33 As a result, MOFs have been designed and developed for various notable applications including heterogeneous catalysis.34 The metalloligand strategy has also been effectively utilized for the design of diverse architectures including MOFs.35–38 A metalloligand allows the incorporation of a desired metal ion that may facilitate light absorption in the visible region of the solar spectrum.39–41 Such a strategy may enable a metalloligand to act as a light harvester in the resultant MOF.35–41 In contrast, use of an organic ligand(s) in a MOF often lacks this feature.42 An efficient photocatalyst is typically composed of a photosensitive unit for light harvesting and an active site for a potential catalytic reaction. Metalloligand-based molecular architectures fulfil both these roles well.35–41 While a metalloligand may act as a light harvester (i.e. photosensitizer), the secondary metal ion(s) may function as the catalytic site.43 Importantly, the electronic structure of the resultant architectures can be manipulated by the selection of (i) the primary metal ion in a metalloligand; (ii) the associated ligand(s) in a metalloligand; and (iii) the secondary metal ion(s).35–38
Only a handful examples of metalloligand-based photocatalytic molecular architectures are available in the literature.44,45 The majority of such examples have utilized photosensitizers that are based on precious metals, such as ruthenium and iridium.43–45 Furthermore, most of these examples have been utilized either for CO2 reduction43 or dye degradation.43–45 Very rarely have such architectures been utilized for visible-light mediated organic transformations.46,47 In these cases, such molecular architectures function as semiconductor photocatalysts.43–45
Typically, in a semiconductor photocatalysis, a catalyst is activated by light, generating energetic electrons and holes.15,16,48 Such species can initiate oxidation and reduction reactions to form a series of reactive oxygen species (ROS), such as a superoxide radical (˙O2−), hydroxyl radical (˙OH) and singlet oxygen (1O2).1–3,48 Together with the photogenerated holes (h+), such ROS can initiate various photocatalytic reactions.4,9–16,48 In well-documented photocatalytic degradations, photogenerated holes and hydroxyl radicals are the major yet non-selective oxidants with the potential to mineralize organic compounds into simpler products such as CO2, N2, H2O, etc.4,37 Similarly, such potent oxidants can also cause non-selective oxidation of organic substrates; although with significantly lower yield and selectivity of the desired product. Selective oxidation can only be accomplished with milder ROS, such as ˙O2− and 1O2.4,9 Therefore, the precise control over the generation of ROS is critical for the desired photocatalytic organic transformation. The selective oxidation of alcohols to aldehydes and ketones is an important organic transformation.49,50 A large number of both natural and synthetic products, pharmaceuticals and drugs contain carbonyl groups, thus emphasizing their significance.49,50 Sulfoxidation is an equally appealing reaction in organic synthesis, with the desulfurization of fossil fuels and disposal of S-based chemical warfare agents being some of the notable categories.51,52
Utilizing semiconductor photocatalysis, there are two pathways for the generation of charge carriers.1–3,48,53,54 The type-I process relies on electron transfer to generate ROS (˙OH, ˙O2− and 1O2).1–5,48 In contrast, the type-II pathway is based on energy transfer from triplet state excitons to molecular oxygen (3O2) to generate the singlet oxygen (1O2).53,54 Besides the lower oxidation potential of a photocatalyst, the formation of triplet state excitons is a prerequisite for the generation of 1O2.53,54 Such a pathway can be achieved either by incorporating heavy atoms (e.g. Bi, I, Se) or by incorporating lanthanide metals that are known for their long triplet state lifetimes.53–55 Taking these critical points into consideration, in this work, a Co3+-based metalloligand 1, offering appended arylacrylate groups, has been utilized for the synthesis of lanthanide-based MOFs (Ln-MOFs), 1-Tb and 1-Eu. Both structurally characterized Ln-MOFs act as semiconductor photocatalysts and support visible light-driven selective oxidation of alcohols and sulfides. Transient absorption and EPR spectroscopy show electron transfer from the light harvesting metalloligands to the Ln3+ ions, generating active Ln2+ sites that function as reaction centers. Photoluminescence and phosphorescence studies validate the intersystem crossing. The triplet state excitons facilitate the exclusive generation of singlet oxygen (1O2), which results in selective oxidation reactions. The photocatalysis results are corroborated by band gap energy calculations and various mechanistic studies, including with a Zn-analogue of 1-Tb that supports the heavy-metal effect of a lanthanide ion in a Ln-MOF. To the best of our understanding, this work not only provides an excellent substrate scope for catalysis but also comprehensive mechanistic details of the entire photocatalytic process.
Fig. 1 Synthetic routes and crystal structures of Ln-MOFs 1-Tb and 1-Eu. Conditions: I = NaH, DMF; II = Co2+ salt, O2; III = OH−; IV = Tb3+ salt; V = Eu3+ salt. |
The presence of large cavities in both Ln-MOFs suggests their porosity, which was substantiated by N2 sorption experiments. The Ln-MOFs, 1-Tb and 1-Eu, exhibited BET surface areas of 42 and 76 m2 g−1, respectively (Fig. S17, ESI†), consistent with their cavity dimensions.57 The PXRD patterns of the as-synthesised samples matched very well with the patterns simulated from the single crystal diffraction analyses, indicating phase purity of the bulk samples (Fig. S18, ESI†). The diffuse reflectance spectra (DRS) of 1-Tb and 1-Eu exhibited absorptions at ca. 620–625 and ca. 425–433 nm (Fig. 2a). The Mott–Schottky plots for both Ln-MOFs show positive slopes at two different frequencies (Fig. 2c and d). This is consistent with the behaviour of a typical n-type semiconductor.58 For 1-Tb, the conduction band position (LUMO), determined from the intersection, was −0.94 V versus NHE. The valence band (HOMO) was then calculated to be 1.31 V versus NHE. Therefore, the band gap energy of 1-Tb was found to be 2.25 eV from the Tauc plot using the direct method (Fig. 2b).59 Similarly, the band gap energy for 1-Eu was found to be 2.30 eV while the LUMO and HOMO potentials were −1.09 and 1.21 V, respectively (Fig. 2b–d). The band gaps for both Ln-MOFs were also fitted using the indirect band gap model as illustrated in Fig. S19, ESI.†,59 The resulting indirect band gaps for both Ln-MOFs were very close (±0.02) to those calculated using the direct model. These parameters suggest adequate light absorption and electron–hole separation ability of both Ln-MOFs upon visible-light irradiation.48 Based on these results, energy profiles of both 1-Tb and 1-Eu are proposed in Fig. 2e. The band gap positions potentially point toward the generation of superoxide (˙O2−) radicals and not hydroxyl (˙OH) radicals upon visible light absorption (vide infra).4,9 These results suggest that the present Ln-MOFs could be utilized for visible-light-driven photocatalysis.
Subsequently, the reaction conditions were optimized taking benzyl alcohol as a model substrate while catalytic reactions were performed in water under visible light irradiation (Table 1). 1-Tb acted as an efficient and selective photocatalyst by only producing benzaldehyde (1a; Scheme 1, vide infra) with >99% isolated yield within 15 min (entry 1, Table 1). This reaction provided a turnover number (TON) of 279 and a turnover frequency (TOF) of 19 min−1. The Eu-analogue, 1-Eu, also resulted in the quantitative oxidation of benzyl alcohol (entry 2). The photocatalytic nature of the oxidation reaction was established by a series of control experiments (Table 1). Reaction without a photocatalyst (entry 3) or in the absence of visible light under otherwise identical conditions (entry 4) did not produce any benzaldehyde. These reactions confirm the photocatalytic nature of the transformation.37 To rule out the possibility of a thermal reaction, catalysis was attempted at elevated temperature (60 °C) but without any light source. Such a reaction only resulted in a negligible amount of a product (entry 5), confirming that the oxidation of benzyl alcohol is exclusively driven by the visible light.49 The photocatalytic reaction was also performed using sunlight as the irradiation source and a nearly quantitative yield was obtained (entry 6). This illustrates that sunlight can be conveniently used for the oxidation reaction. As expected, no product was produced in the absence of O2 (entry 7), while the presence of air reduced the benzaldehyde yield (entry 8). Importantly, metalloligand 1 did not promote the photocatalytic oxidation of benzyl alcohol (entry 9; vide infra). The 1-Tb catalyzed oxidation of benzyl alcohol was used to investigate the reaction kinetics.16 A plot of ln(Ct/Co) versus time revealed first-order kinetics with a rate constant of 9.27 × 10−2 min−1 (Fig. 3a and S20, ESI†).37 A high-rate constant confirms a fast photocatalytic reaction. It can be noted that rate constants have rarely been calculated for photocatalytic reactions;15,16 while rate constants using external oxidants (e.g. H2O2) are inferior to that of the present Ln-MOFs.4,5,15,16
Entrya | Reaction parameters | Isolated yield (%) |
---|---|---|
a Reaction conditions: benzyl alcohol (0.92 mmol), photocatalyst (1 mol%), water (2 mL), temperature (25 °C). TON = molbenzaldehyde × mol1-Tb−1; TOF = TON/15 min. NR = No reaction. | ||
1 | 1-Tb, O2, 15 min | >99 |
TON = 279 | ||
TOF = 19 min−1 | ||
2 | 1-Eu, O2, 20 min | >99 |
3 | No catalyst, O2, 60 min | NR |
4 | 1-Tb, O2, no light, 60 min | NR |
5 | 1-Tb, O2, 60 °C, no light, 30 min | 6 |
6 | 1-Tb, O2, sunlight, 30 min | 98 |
7 | 1-Tb, N2, 30 min | NR |
8 | 1-Tb, air, 15 min | 68 |
9 | 1, O2, 30 min | NR |
10 | 1-Tb, O2, NaN3, 15 min | 12 |
11 | 1-Tb, O2, isopropyl alcohol, 15 min | 98 |
12 | 1-Tb, O2, p-benzoquinone, 15 min | 84 |
Scheme 1 Substrate scope for the photocatalytic oxidation of assorted alcohols and a scale-up reaction. Reaction conditions: substrate (0.92 mmol), 1-Tb (1 mol%), H2O (2 mL), temperature (25 °C). All yields are for the isolated products. The characterization data and the corresponding spectra for the representative products are presented in Fig. S22–S84, ESI.† |
To investigate the nature of the ROS involved in the oxidation of benzyl alcohol, different scavengers were used. The use of NaN3, isopropyl alcohol and para-benzoquinone are known to specifically inhibit the generation of 1O2, ˙OH and ˙O2−, respectively.37,60 The oxidation of benzyl alcohol was significantly suppressed by the use of NaN3 (<12% isolated yield, 88% inhibition, entry 10). Meanwhile, the presence of isopropyl alcohol (.OH scavenger) did not affect the oxidation of benzyl alcohol (entry 11), and the use of para-benzoquinone (˙O2− scavenger) only partially inhibited its oxidation (entry 12). These results suggest that 1O2 is the primary ROS involved in the photocatalytic oxidation of benzyl alcohol (vide infra).
To further confirm the generation of 1O2 by light irradiation, a few 1O2 selective probes, such as 9,10-dimethylanthracene (DMA) and 1,5-dihydroxynapthalene (DHN) were utilized.49,61 The intense fluorescence of DMA, at 427 nm, was significantly quenched in the presence of 1-Tb under visible light irradiation. This indicates that DMA was oxidized by 1O2 to endoperoxide, DMA-O2 (Fig. 3b).49 Likewise, the absorbance of DHN at 301 nm was decreased upon exposure to visible light in the presence of 1-Tb. Furthermore, a new band was generated at 420 nm due to the photooxidative depletion of DHN to form 5-hydroxy-1,4-napthalenedione (juglone) on reaction with 1O2 (Fig. 3c).61 These results corroborate the generation of 1O2 as the predominant ROS. The kinetics were studied by monitoring the change in the absorption peak of juglone at 420 nm, and the rate constant was calculated to be 0.0542 min−1 (Fig. S21, ESI†). The calculated rate constant is higher than that observed for other common photocatalysts reported in the literature (Table S6, ESI†).61 This indicates superior capacity of the present Ln-MOFs for generating 1O2. Finally, a very high concentration of 1O2, based on 1-Tb was calculated to be ca. 84 μM.
Since radical scavenger experiments suggest that species other than 1O2 could also be involved in substrate oxidation, we investigated whether other ROS (˙OH and ˙O2−) are also formed in good quantities. Coumarin is particularly helpful for identifying ˙OH radicals as it forms 7-hydroxycoumarin, which is emission active.37,62 However, insignificant emission enhancement was noted, suggesting that the hydroxyl radicals are not the primary oxidant in the present case (Fig. 3d). This is further supported by the fact that the HOMO potentials for both Ln-MOFs are not sufficiently positive (EVB: 1-Tb = 1.31 V, 1-Eu = 1.21 V) to support the generation of a highly active yet non-selective oxidant, ˙OH [E(˙OH/OH−) = 1.89 V vs. NHE] (Fig. 2e).9
The scavenger nitro-blue tetrazolium (NBT) is exclusively selective for ˙O2−.37,62 During the course of reaction for 60 min in an O2 saturated environment, a prominent absorption band was observed at 530 nm due to the formation of NBT-formazan. A spectral titration allowed the calculation of the concentration of ˙O2− radicals, which was found to be ca 6 μM (Fig. 3e, inset). Importantly, however, the concentration of NBT-formazan species decreased rapidly. This suggests two possibilities; either consumption of ˙O2− radicals or conversion of ˙O2− radicals into some other species (Fig. 3e). The consumption of ˙O2− radicals is ruled out in the absence of any suitable reactant. However, ˙O2− radicals are known to react with holes (h+) to form 1O2.4,62 In fact, the generation of 1O2 is certainly conceivable in the present case as per the energy profile (cf.Fig. 2e).
In essence, the proposed mechanism for the oxidation of benzyl alcohol is consistent with the activation of O2 to primarily produce 1O2 (ca. 84 μM) along with the limited generation of ˙O2− (ca. 6 μM) under visible light irradiation. The band gap energies for both Ln-MOFs (1-Tb: 2.25 eV, 1-Eu: 2.30 eV) are ideal, favouring a facile generation of the mild oxidant, 1O2, while inhibiting the generation of a stronger but non-selective oxidant, ˙OH. To further provide evidence for the generation of ROS, electron paramagnetic resonance (EPR) spectra were recorded under visible light irradiation. Upon visible light irradiation, a sharp EPR spectral signal was detected at g = 2.004 (ΔH = 4 G), supporting a radical-based-intermediate (1O2/˙O2−) (Fig. 3f).62
Of the two MOFs, 1-Tb catalyzed the oxidation of benzyl alcohol in a shorter span of time, therefore acting as a better photocatalyst (entries 1 and 2, Table 1). Therefore, 1-Tb was used to explore the substrate scope by attempting the oxidation of assorted alcohols. Various mono- (1b–1l), di- (1m) and even tri-substituted (1n) benzyl alcohols were quantitatively oxidized with exclusive selectivity for an aldehyde (Scheme 1). The type and nature of a substituent(s) on an aromatic ring did not play a significant role in the reaction outcome, since all corresponding products were obtained in nearly quantitative yields. These results strongly suggest a radical mediated catalytic pathway. However, substrates bearing electron-donating groups required a slightly longer time for the completion of a reaction when compared to the substrates with electron-withdrawing groups.49 Similarly, o-substituted benzyl alcohol (1i) needed a longer duration than that of m- and p-isomers (1j and 1k), most probably due to the steric hindrance.63
Substrates bearing multiple aromatic rings such as 4-biphenylmethanol and 2-napthylmethanol were successfully oxidized providing the corresponding aromatic aldehydes in excellent yields (1o, 1p). Likewise, dihydroxy substrates, such as hydrobenzoin and 1,3-phenylenedimethanol, were also efficiently oxidized to their corresponding carbonyl compounds in quantitative yields (1q, 1r). On photocatalytic oxidation of cinnamyl alcohol (allylic alcohol), the desired α,β-unsaturated aldehyde was obtained in 96% yield (1s), thus highlighting the chemoselectivity.63 Notably, aliphatic alcohols, such as undecane-1-ol and morpholinomethanol, also afforded the desired aldehydes (1t, 1u) in high yield and exclusive selectivity. These examples justify the present Ln-MOFs as effective photocatalysts for the oxidation of aliphatic alcohols. It is significant that the selective oxidation of aliphatic alcohols has seldom been reported as controlling their over-oxidation is particularly challenging.64
In order to further expand the substrate scope, the oxidation of secondary aliphatic alcohols was attempted. For example, both 1,7,7-trimethylbicyclo[2.2.1]heptan-2-yl)methanol and octan-2-ol yielded the corresponding ketones in high yields (1v, 1w). The oxidation of heteroaromatic alcohols such as 3-pyridinemethanol, furan-2-ylmethanol, 5-nitrofuran-2-ylmethanol and (5-(3-chloro-4-methoxyphenyl)-furan-2-yl)methanol also afforded the expected aldehydes in excellent yields (1x-1aa). Collectively, a broad range of both primary and secondary alcohols were exclusively oxidized to their corresponding aldehydes and ketones. In order to illustrate the oxidation of challenging substrates, a few steroid-based aliphatic alcohols such as 20α-hydroxysteroid progesterone, 17β-estradiol and 3α/β-androstanediol were employed.63 Notably, in all three cases, the corresponding products were obtained in excellent yields (1ab-1ad). Finally, considering a straightforward process, a gram scale reaction was performed. With 1 g of benzyl alcohol, the reaction took longer (2 h), but an exclusive and quantitative amount of benzaldehyde was obtained (0.95 g; 97%), thus demonstrating a remarkable catalytic performance of Ln-MOFs (Scheme 1).
Although a few reports in the literature mention the use of nanoparticles and composite-materials for the selective oxidation of alcohols,49 very few MOFs are known to be successful for this reaction.55,65 Importantly, the photocatalytic efficiency of the present Ln-MOFs for alcohol oxidation is far superior to that of most of the previously reported MOFs and composite materials (Table S7, ESI†). More importantly, the reported examples lack a detailed substrate scope and mechanistic investigations (vide infra).49,55,56,65
In the present catalytic system, the oxidation of an alcohol is predominantly associated with the action of 1O2, as suggested by the scavenger studies. Based on these findings, a plausible mechanism is proposed in Fig. S85, ESI.† Upon visible light irradiation, a Ln-MOF gets photoexcited followed by the separation of an electron and a hole. The photo-electron activates O2 to 1O2 which oxidizes benzyl alcohol to a benzyloxy radical (I) that rearranges to a carbon-centred radical (II).65 The species II subsequently reacts with ˙OOH to form a peroxy hemiacetal (III) which decomposes to give benzaldehyde as the final product and hydrogen peroxide (H2O2) as a by-product. Importantly, the by-product H2O2 was confirmed using UV-Vis spectral monitoring of triiodide ions using an aqueous solution of KI (Fig. S86, ESI†).66 More importantly, H2O2 was quantitively estimated by titrating it against a KMnO4 solution that provided its concentration as 2.68 μM.67 Considering that H2O2 could also be a potential oxidant, a control experiment involving gallic acid as an H2O2 scavenger was carried out.68 Notably, however, the quantitative oxidation of benzyl alcohol to benzaldehyde was noted in the presence of gallic acid. This confirms that the photocatalytic oxidation of benzyl alcohol does not depend on in situ formed H2O2.
Scheme 2 Substrate scope for the photocatalytic oxidation of assorted sulfides, a scale-up reaction and photocatalytic oxidation of 2-chloroethyl-ethyl-sulfide (CEES). Reaction conditions: substrate (0.80 mmol), 1-Tb (1 mol%), MeOH (2 mL), temperature (25 °C). All yields are for the isolated products. The characterization data and the corresponding spectra for representative products are presented in Fig. S91–S113, ESI.† |
1-Tb exhibited a good substrate scope with very high yield for most of the sulfides. Photooxidation of diphenylsulfide, dibenzylsulfide and their substituted analogues bearing both electron-withdrawing and electron-donating groups produced the corresponding sulfoxides in excellent yields (2b–2f). A nearly quantitative yield was obtained for the photooxidation of tetrahydrothiophene (2g), a cyclic sulfide. Importantly, several aliphatic sulfides (n = 4, 8 and 10) also yielded the respective sulfoxides (2h–2j) in high yields. More importantly, both the yield and the selectivity were not compromised for any such substrate. Subsequently, sulfoxidation of thioanisole was carried out on a gram scale that exclusively yielded 1.02 g of methylphenyl sulfoxide in 92% isolated yield (Scheme 2). The same reaction was used to monitor the rate of sulfoxidation. The formation of methylphenyl sulfoxide was found to follow first-order kinetics with a rate constant of 0.0562 min−1 (Fig. 4a and S114, ESI†). The photooxidation efficiency and selectivity of 1-Tb towards different sulfides either matched or exceeded the reported examples.51,54,60 It is worth mentioning that several such literature examples suffered from the formation of undesirable sulfone (Table S9, ESI†).60
Finally, 1-Tb was used for the detoxification of a mustard gas simulant, 2-chloroethyl-ethyl-sulfide (CEES).51,52 Only 1 mol% of 1-Tb exclusively oxidized CEES to non-toxic 2-chloroethyl-ethyl-sulfoxide (CEESO, 2k). The selectivity for this photooxidation was evaluated by proton NMR spectral experiment (Fig. 4b). New signals were identified at 1.37, 2.85, 2.95, 3.19, and 3.98 ppm, which were consistent with the formation of CEESO replacing the characteristic CEES peaks at 1.27, 2.61, 2.88, and 3.67 ppm. The reaction was complete within 15 min and only sulfoxide was formed as the exclusive product. Even after 24 h of the initial photo-irradiation, no sulfone could be detected. Similar to the selective oxidation of alcohols, only a few materials are known for the oxidation of sulfides.51,52 In comparison to the reported examples, the present Ln-MOF photocatalysts offered a considerable improvement in the photocatalytic activity, selectivity and substrate scope.61,65
Based on radical scavenger experiments, 1O2 is the key ROS involved in the oxidation reactions. Accordingly, a mechanistic pathway for the photooxidation of CEES is proposed (Fig. 4c). Upon nucleophilic addition of the thioether substrate to the electrophilic 1O2, a weakly bound persulfoxide intermediate I is formed. Intermediate I undergoes quenching in a polar aprotic solvent while it is stabilized in a protic solvent by H-bonding, thus substantially enhancing the effectiveness of the photooxidation in polar protic solvents.51,52 Control experiments also supported this assertion (Table S6, ESI†). As a consequence of the nucleophilic reaction between intermediate I and another molecule of CEES, an unstable anionic hypervalent intermediate II is formed.51,52 Such an unstable intermediate breaks down to give two molecules of sulfoxide, CEESO. Therefore, each molecule of singlet oxygen has the ability to produce two equivalents of sulfoxide, justifying a fast and efficient oxidation of CEES to CEESO.
For 1-Tb, along with a 5 nm blue shift of λESAmax to 560 nm, unlike the TA spectra of both L and 1, the growth of a shoulder in the ESA is seen to develop at around 200 fs (with the wavelength maximum of the shoulder being at around 470 nm), and this develops into a negative ΔA peak at around 450 nm at 20 ps (Fig. 5c). Moreover, the global bi-exponentially fitted kinetic profile also reveals an interesting probe wavelength dependency (Fig. S115, ESI†). For the probe wavelength range of 450–535 nm, an increase in the signal (with negative amplitude) of ∼2 ps is observed alongside a slower component of ∼6 ps (IC; τD = time scale of intra-molecular vibrational relaxation; Fig. S115c, ESI†). Beyond 540 nm, this growth component is replaced by a decay component with a long lifetime of 2075 ps (Fig. S115c, ESI†). We propose that this increase in signal, which implies the development of a photoproduct, corresponds to the generation of a Tb2+ state after a facile electron transfer from the metalloligand to the Tb3+ ion on a timescale of ∼2 ps (τET = time scale of electron transfer).
To gain further insights into the photo-induced electron transfer, EPR spectral experiments were carried out.13 The EPR spectrum of 1, under dark conditions and a N2 environment, did not exhibit any signal owing to the presence of a Co3+ ion (d6, S = 0). Upon visible-light irradiation, a distinct signal at g = 2.01 (ΔH = 6 G) was observed (Fig. 5g). When the light was turned off, the signal immediately disappeared, thus confirming visible-light-induced radical formation. A relatively small value of g suggests that electron transfer during the photo-induced process did not occur from the primary metal ion, Co3+, but rather from the organic ligand of metalloligand 1. To substantiate this, L was also subjected to the EPR spectral investigation under visible light-irradiation. The EPR spectrum of photo-induced L (g = 2.02; ΔH = 4 G) was very weak but similar to the one observed for 1 (Fig. 5h). This confirms that photo-induced electron transfer occurs from the organic ligand of metalloligand 1. Importantly, the EPR spectrum of 1-Tb, under an inert atmosphere, showed a broad signal at g = 2.00 (ΔH = 8 G), corresponding to a Tb3+ ion.70 Upon visible light irradiation, a rather complex EPR spectrum was obtained (Fig. 5i), in which a sharper signal is attributed to the paramagnetic Tb2+ ion (g = 1.95; ΔH = 4 G). These results unambiguously validate electron transfer from a metalloligand to a Tb3+ ion thus generating a Tb2+ site. To further support this, EPR spectra were also measured for 1-Eu. In the absence of visible light, no signal was detected as both Co3+ and Eu3+ ions are EPR inactive.39 However, upon visible light irradiation, signals at g = 2.05 (ΔH = 4 G) and g = 4.20 (ΔH = 9 G) appeared (Fig. 5j). Such an EPR spectrum is attributed to a paramagnetic Eu2+ species, generated as a result of electron transfer from a metalloligand to an Eu3+ ion.39 Subsequently, an excited Ln2+ ion transfers its energy to the triplet oxygen (3O2) to generate singlet oxygen (1O2), which is responsible for the photooxidation reactions. Finally, electron transfer from a Ln2+ ion to a metalloligand completes the catalytic cycle. In summary, Fig. 5k illustrates a schematic representation of ultrafast TAS results for L, 1 and 1-Tb.
To further support electron transfer, the photoluminescence (PL) of metalloligand 1 and Ln-MOF 1-Tb were analysed. Both 1 and 1-Tb displayed strong absorption bands at 400–600 and 400–650 nm, respectively (inset of Fig. 6a and 2a). While an intense absorption band at 400–500 nm corresponded to a Co3+-centered metal-to-ligand charge transfer (MLCT) transition, a relatively weak absorption band, at higher wavelength, is ascribed to the d–d transition of the primary Co3+ ion. Notably, the absorption spectrum of 1-Tb closely matches that of metalloligand 1, thus suggesting that the coordination of the secondary metal ions (Tb3+) has not significantly perturbed the electronic environment of the Co3+ ion in 1.24 It was thus assumed that coordination of a Tb3+ ion to a metalloligand should not significantly affect its PL properties. However, the PL intensity of 1-Tb (at 452 nm) was significantly quenched when compared to the metalloligand 1 (Fig. 6a). A much lower emission intensity for 1-Tb suggests a reduced population of the singlet excitons since fluorescence results from the radiative decay of the singlet excitons.55 Notably, typical Tb3+-based f–f emissions were also not observed in the PL spectrum of 1-Tb. This further confirms electron transfer from a metalloligand to a Tb3+ ion in the resultant Ln-MOF.71,72 Additionally, it suggests that the recombination rate of photo-generated charge carriers is higher in metalloligand 1 than in the Ln-coordinated-metalloligand (i.e.1-Tb).71 This is further validated by EIS spectroscopy.72 The superior charge separation effectiveness of 1-Tb over metalloligand 1 was confirmed by its lower arc radius in the Nyquist plot (Fig. 6b). These results convincingly support the assertion that 1-Tb is a highly efficient photocatalyst when compared to metalloligand 1. The catalysis results proved the same.
Furthermore, from decay associated spectra of 1-Tb (DAS; Fig. S117 and S118, ESI†), it is again evident that a decrease in the amplitude of the e-transfer component coincides well with an increase in the amplitude of the ∼2 ns long decay component. This implies that the disappearance of the Tb2+ state is accompanied by the appearance of this long component, which is attributed to inter-system crossing (ISC) induced by the reduced Tb2+ ion (τISC = time scale of inter-system crossing).51,52,61 Thus, singlet excitons might have transitioned to triplet excitons, as illustrated by the TAS studies.51,52,61 To understand this phenomenon better, phosphorescence spectra of 1 and 1-Tb were measured and compared.55 Ln-MOF 1-Tb exhibited a strong phosphorescence peak at 519 nm (Fig. 6c), displaying a 67 nm red-shift with respect to the fluorescence peak at 452 nm. In fact, a longer lifetime of 1-Tb (2.01 ms) further shows that the red-shifted emission is due to the triplet excitons (Fig. 6d).55 In contrast, metalloligand 1 displayed a negligible intensity in its phosphorescence spectrum. A significant difference in the phosphorescence intensities of 1-Tb and 1 (lifetime = 8.29 μs; Fig. 6d, inset) suggests that 1-Tb generates triplet excitons that are not produced by the metalloligand alone. Hence, it is evident that the generation of the Tb2+ state facilitates singlet-to-triplet conversion followed by emission via phosphorescence to the singlet ground state. One should also take note of the fact that the amplitude of the rise component starts decreasing (i.e. becoming less negative) from around 500 nm and is lowest at 530 nm. This neatly coincides with the maximum of the phosphorescence spectrum and concomitant appearance of the ∼2 ns ISC component. In summary, Fig. 6e presents a schematic representation of the ISC exhibited by Ln-MOF 1-Tb.
The newly synthesized 1-Zn provided an excellent opportunity to test the heavy-atom effect of Tb3+ ions in 1-Tb in generating the triplet excitons. Similar TA spectral experiments were carried out with 1-Zn and the λESAmax was noted to be 575 nm (Fig. 7f and g). The global analyses revealed the presence of two decay components of 7.5 ps (IC) and 53.3 ps (excited state decay), while no probe-dependent rise component was observed over the entire wavelength range (Fig. S124, ESI†). Therefore, the absence of both growth phase and long decay components suggests the absence of electron transfer and heavy-metal induced ISC, and the latter indicates the absence of phosphorescence.
Gratifyingly, the absence of a signal in the phosphorescence spectrum of 1-Zn (Fig. 7h) indicates that in the absence of lanthanide ions, triplet exciton generation is not a favorable pathway. In fact, non-selective oxidation of benzyl alcohol to benzoic acid by using 1-Zn pointed to the generation of strongly-oxidizing ˙OH radicals. This was proved by the radical scavenger experiments involving isopropyl alcohol, a known scavenger for ˙OH radicals, which completely inhibited the oxidation of benzyl alcohol.9–14,25–28 Finally, the presence of ˙OH radicals was confirmed by the selective oxidation of coumarin to 7-hydroxycoumarin, monitored by emission spectroscopy (Fig. S125 and S126, ESI†), which provided its concentration as ca. 10 μM.
In essence, it is proposed that upon photoexcitation, electron transfer (Table S10, ESI†) takes place from the Co3+-based metalloligands (1; τav(1) = 24.50 ps) to the secondary Tb3+ ions in 1-Tb (τav(1-Tb) = 1034.60 ps) generating the active Tb2+ sites which are responsible for the generation of triplet state excitons that activate triplet oxygen (3O2) to singlet oxygen (1O2). The singlet oxygen, thus generated, selectively catalyses the oxidation reactions. While evidence of a Tb2+ ion was obtained from the EPR spectrum, phosphorescence of 1-Tb with a lifetime of ∼2 ms further confirms ISC from the excited first singlet state (S1) to the triplet manifold of the metalloligand induced by the heavy Tb atom. However, only ligand L and only metalloligand 1 solely cannot induce the generation of triplet excitons as evident from both TAS and phosphorescence studies. Therefore, both L and 1 are not capable of catalyzing any visible-light mediated catalytic reaction. Similarly, the non-heavy-metal analogue of 1-Tb, i.e.1-Zn (τav(1-Zn) = 38.06 ps), cannot facilitate ISC that favours the generation of triplet excitons. In conclusion, a Co3+-based metalloligand 1 functions as a photosensitizer to facilitate light absorption in the visible region while the catalytic reactions take place at the secondary Ln sites in the resultant Ln-MOFs.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2288468 and 2288469. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4sc04105d |
This journal is © The Royal Society of Chemistry 2024 |