Subhrajyoti
Ghosh‡
a,
Srijan
Mukherjee‡
a,
Veerappan
Karthik
b,
Priti
Bera
a,
Amarajothi
Dhakshinamoorthy
*bc and
Shyam
Biswas
*a
aDepartment of Chemistry, Indian Institute of Technology Guwahati, Guwahati, 781039, Assam, India. E-mail: sbiswas@iitg.ac.in
bSchool of Chemistry, Madurai Kamaraj University, Madurai, Tamil Nadu 625021, India
cDepartamento de Química, Universitat Politècnica de València, Camino de Vera s/n, Valencia 46022, Spain. E-mail: admguru@gmail.com
First published on 20th February 2024
Modern-day environmental pollution does not end with only air pollution by harmful gases. Issues also arise due to dust particles released by industry and motor vehicles. Regular use of herbicides, fungicides, and pesticides to increase the production of foodstuffs in agricultural fields also enhances the risk of soil and aquatic environmental pollution. Therefore, the systematic development of multi-functional materials for the catalytic conversion of toxic gases and hydrocarbons into value-added products under mild conditions, selective detection, and quantification of toxic organic pollutants in environmental water bodies are crucial to maintaining a sustainable environment. Herein, we have strategically engineered a highly robust, fluorine-rich, Zr(IV)–organic framework (1) and, desolvated (1′) to address three emerging environmental issues. This is the first recyclable metal–organic framework (MOF) catalyst (1′) where the Lewis basic nature of fluorine atoms and the Lewis acidic nature of the Zr(IV) sites are utilized for the mild pressure, solvent-free entrapping of CO2 to epoxides with various kinetic dimensions in high yields. The hydrophobic cavity of 1′ also facilitated better conversions of hydrocarbons into their respective alcohols/ketones compared to non-hydrophobic MOF under aerobic oxidation conditions with recyclability. Notably, for the first time, selective, rapid (response time <5 s) and nanomolar (LOD = 1.1 nM) fluorescence-based detection of aclonifen (ACF) herbicide was performed using 1′ in various environmental water samples, at different pH conditions, and also in variety of food and vegetable extracts. Moreover, a reusable 1′@starch@cotton composite was fabricated for the on-site, naked-eye detection of ACF under UV light. Additionally, a detailed investigation of the mechanistic pathways of all the proposed applications was performed experimentally and theoretically.
O bond enthalpy (805 kJ mol−1).3 Various chemical fixation strategies4 have been proposed, including entrapping CO2 in an epoxide ring to produce cyclic carbonate. This process gained huge importance due to the potential application of cyclic carbonates as fuel additives and electrolytes.5 Moreover, various homogeneous catalysts have been reported for selective conversion of CO2 into cyclic carbonates with 100% atom economy.6 Despite all these advantages, the poor separation efficiency of the product and the lack of recyclability make these processes less economical. On the other hand, various MOF-based heterogeneous catalysts have gained considerable attention for the catalytic conversion of CO2.3 The highly porous nature, as well as the presence of three-dimensional channels and Lewis acidic and basic functionalities in MOFs offer several advantages for the selective adsorption and activation of CO2.2,3,7 However, very few MOFs have been reported for conversion of CO2 under mild pressure and temperature conditions; in most of the reported catalysts, the lone pair of ‘N’ atom acts as a Lewis base.7,8 In this regard, for the first time, we have demonstrated the usefulness of fluorine-functionalized highly porous, superhydrophobic desolvated Zr(IV)-MOF (1′) for selective chemical fixation of CO2. The presence of a strongly electron-dense pentafluorobenzene functionality provides two advantages. Firstly, the high fluorine content of the linker enhances the catalyst's hydrophobicity and aqueous/moisture phase stability. Secondly, highly electron-dense fluorine can act as a Lewis basic site for selective interaction and adsorption of CO2 on the porous surface. Lewis basic activity of the fluorine and its influence on the catalytic conversion of CO2 in cyclic carbonates is less explored.9,10 Hence, scientific efforts should be made in this direction.
The presence of a highly electronegative –F group in the linker moiety may be helpful for selective hydrophobic–hydrophobic interaction with hydrocarbons. Hence, this advantage of hydrophobicity in 1′ was further explored in the catalytic aerobic oxidation of hydrocarbons to their corresponding alcohol (ol)/ketone (one) mixture with high selectivity (>90%), which is one of the important reactions in organic chemistry and fine chemical industry.11 To achieve this target, a series of heterogeneous catalysts have been reported in which hydrocarbons are converted to their respective ol/one mixture. Among the various heterogeneous catalysts, MOF-based heterogeneous solids have received considerable attention due to the possibility of the preparation of MOF catalysts with different transition metals or active sites encapsulated within the pores of MOFs.12–15 Among these different MOF catalysts, UiO-66-based solids have shown superior performance due to the presence of defective sites within the framework.16 Besides these conventional MOF catalysts, porphyrin-based MOFs have also been reported as oxidation catalysts.17 Due to these features of MOFs compared to other heterogeneous catalysts, wide ranges of MOFs have been reported for the aerobic oxidation of hydrocarbons, including Au-Pd/MIL-101,18 MIL-101,19 NHPI/Fe(BTC),20 Fe(BTC)/TBAB,21 {[Cu0.5La2(HPDC)(PDC)2(SO4)(H2O)2]H2O}n,22 NHPI/Fe(BTC),23 Cu-doped ZIF-824 and Ce-based MOF.25 However, the catalytic performance of a MOF solid in the oxidation of hydrocarbon depends on many factors, such as the nature of active sites, the nature of the oxidant, reaction temperature, and solvent.14,26 Although various types of MOFs are utilized to catalyse this reaction, very few of them are hydrophobic in nature.
The introduction of hydrophobic groups (–F atoms) in the organic linker of a MOF permits the accommodation of hydrophobic sites, which can be responsible for increasing the concentration of hydrocarbons within the framework of the MOF by providing a confined environment. This further provides closer proximity among the active sites of the oxidant, and the hydrocarbon to initiate a catalytic oxidation reaction. For instance, in a previous study, PCN-222(Fe)-F7 was employed as a heterogeneous hydrophobic MOF catalyst for the oxidation of cyclohexane and showed 50% conversion with 90% selectivity to ol/one mixture using t-butylhydroperoxide (TBHP) at 80 °C in acetonitrile.27 In another report, PCN-224(Mn) MOF was exchanged with 2,2′-bis-(trifluoromethyl)-4,4′-diphenyl phthalate (L) to create hydrophobic environment in PCN-224(Mn)-L. The catalytic performance of PCN-224(Mn)-L in the aerobic oxidation of cyclohexane was observed with 51% conversion and 90% selectivity to ol/one using TBHP as the oxidant at 80 °C in acetonitrile.28 In contrast, the conversion of cyclohexane in the presence of PCN-224(Mn) presented only 16% conversion with 79% selectivity to ol/one under similar conditions. Recently, we have reported that the conversion of cyclohexane to its respective ol/one mixture is significantly higher (21%) in the presence of DUT-52 MOF with trifluoroacetamido-functionalized linker in acetonitrile at 60 °C after 24 h, while the conversion of cyclohexane with DUT-52 alone was only 6% under similar conditions.29 This lower activity in the absence of a hydrophobic environment is due to the poor interaction between the redox sites and cyclohexane, as cyclohexane remains at the exterior surface most of the time. In this study, by using superhydrophobic 1′, under mild reaction conditions, we achieved the maximum conversion of 69% and 100% selectivity for some hydrocarbons. Compared to these previous attempts, a superior activity was observed in the oxidation of cyclohexane with a high ol/one ratio and higher turnover number as a function of the framework pore diameter near the iron site in Fe2(dobdc) (dobdc4− = 2,5-dioxido-1,4-benzenedicarboxylate) MOF.30 In another work, the selective oxidation of methane to CH3OH was effectively altered by tuning the microenvironment through the hydrophobic modification around Fe-porphyrin within the MOF framework.31 Furthermore, a highly selective oxidation of methane to CH3OH was reported by surface hydrophobic modification over Cu-BTC MOF at low temperature with a yield of 10.67 mmol gcat−1 h−1 and 99.6% selectivity.32 These pertinent precedents clearly illustrate the important role of hydrophobic environments within the MOF structure to alter the reactivity and product selectivity compared to non-hydrophobic MOFs.
Aclonifen (ACF) is a herbicide of the diphenyl ether family often used in agricultural fields to control weeds in various crops.33 It is predominantly used to control grass and broadleaf weeds around crops, such as peanuts, soybeans, and sunflowers.33 ACF selectively interferes with the growth and development of unwanted grass and broadleaf weeds by disrupting their photosynthetic ability or by inhibiting certain metabolic processes.34 This selective action allows crops to thrive, while minimizing the impact on non-target plants.35 The widespread use of this herbicide all over the world for improvement of the efficiency of agricultural production increases the risk of environmental pollution.35 Rural water bodies are mainly contaminated by the extensive use of ACF. Endocrine disruption, potential teratogenicity, limited mutagenic activity, and even some uncertain suspected environmental impacts are few of the hostile effects of ACF. Therefore, extensive exposure to ACF may adversely impact aquatic ecosystems, as well as human health. Hence, regular vigilance regarding the concentration of such toxic herbicides is crucial in natural aquatic systems. Therefore, the development of an economical, reliable sensor for fast, precise, ultrasensitive probing of ACF is required.
Collectively, based on these three points, we have successfully employed 1′ to promote two important organic reactions; (i) mild pressure cycloaddition of CO2 with epoxides and (ii) aerobic oxidation of hydrocarbons under mild reaction conditions. Moreover, ultrasensitive, selective probing of ACF by 1′ in an aqueous medium is also explored. Some of the notable novelties of 1′ are the lowest ever reported limit of detection (LOD) value (1.1 nM), fast response time (<5 s), selective recognition of ACF in water, high sensitivity in wide pH conditions, and real field detection of ACF from various fruits and vegetable extracts. Additionally, fabrication of 1′ over the surface of starch@cotton bio-composite is also demonstrated for the on-site, naked eye nanomolar level detection of ACF under UV light. Further, a detailed mechanistic investigation of all the applications is executed both experimentally and theoretically.
![]() | ||
| Fig. 1 (a) Pawley plot for the PXRD profile of 1′. (b) Cubic framework structure of 1′. (c) and (d) Structures of the free linker and ACF herbicide. | ||
Along with chemical stability, determining the thermal stability of 1′ is an important experiment. Thus, we executed the TG analysis under an O2 atmosphere at the temperature range of 30–700 °C with a continuous heating rate of 4 °C min−1. From Fig. S11 (ESI†), it can be concluded that both 1 and 1′ were thermally stable up to 300 °C. In the TG analysis curve of 1, initially, in between 30–125 °C, 3.1% weight loss occurs due to the removal of 5 water molecules per formula unit. Within the temperature range of 125–300 °C, the loss of 6 DMF molecules resulted in a 12.5% loss of weight. At last, after 300 °C, the collapse of 1 begins because of the cleavage of the Zr–O bonds which resulted in the final weight loss from 1. The single-step weight loss after 300 °C was perceived in the case of 1′ due to the breakdown of Zr–O bonds of the MOF.
| Entry | Catalyst | Co-catalyst (loading) | Temperature (°C) | Conversionb (%) |
|---|---|---|---|---|
| a Reaction conditions: epichlorohydrin (20 mmol), catalyst (15 mg), TBAB (in mol%), CO2 pressure (1 bar), 24 h. b Determined by 1H-NMR. | ||||
| 1 | 1′ | — | 80 | — |
| 2 | 1′ | TBAB (4%) | 80 | >99 |
| 3 | 1′ | TBAB (1%) | 80 | >99 |
| 4 | 1′ | TBAB (0.8%) | 80 | >99 |
| 5 | 1′ | TBAB (0.4%) | 80 | 98 |
| 6 | 1′ | TBAB (0.1%) | 80 | 98 |
| 7 | — | TBAB (0.4%) | 80 | 52 |
| 8 | 1′ | TBAB (0.4%) | 60 | 85 |
| 9 | 1′ | TBAB (0.4%) | 25 | 27 |
| 10 | ZrCl4 | — | 80 | — |
| 11 | H2L linker | — | 80 | — |
| 12 | ZrCl4 | TBAB (0.4%) | 80 | 68 |
| 13 | H2L linker | TBAB (0.4%) | 80 | 49 |
| 14 | ZrCl4 + H2L linker | TBAB (0.4%) | 80 | 66 |
![]() | ||
| Fig. 2 Time conversion plot for the cycloaddition of CO2 with epichlorohydrin employing 1′ under optimized conditions (entry 5, Table 1). | ||
The high catalytic activity of 1′ inspired us to further explore its catalytic activity under optimized conditions with other epoxides possessing different substituents in the side chain. Experimental observations showed that the conversion efficiency almost remained unaltered with smaller epoxides (epichlorohydrin and epibromohydrin) (Table 2, entries 1, 2 and Fig. S21, S31, ESI†). These catalytic results reveal that the pore size of 1′ fits perfectly with the size of the substrate. Moreover, this enables easy adsorption, diffusion, and desorption of the product from the porous catalyst. Hence, a higher conversion rate was obtained. With the increase in the alkyl side chain length, the percentage conversion of epoxides decreased (Table 2, entries 3–7 and Fig. S32–S36, ESI†). For instance, 1,2-epoxydodecane has a larger size compared to the pore of 1′, and diffusion of the substrate to the active site is strongly hindered by the bulky side chain. Hence, the rate of conversion also decreased. Interestingly, 100% selectivity to cyclic carbonate products was observed for all the substrates tested. Further, we also checked the catalytic efficiency of unfunctionalized and amino-functionalized MOFs under similar optimized conditions (Fig. S38–S39, ESI†). A comparatively lower rate of conversion was obtained. Hence, active participation of Lewis basic –F functionality in the catalytic cycle could be established. The cumulative results indicate that 1′ can act as a highly stable and active heterogeneous catalyst for ambient pressure solvent-free fixation of CO2.
| Entry | Catalyst | Co-catalyst | Epoxide | Conversion (%) |
|---|---|---|---|---|
| a Reaction conditions: epoxide (20 mmol), 1′ (15 mg), TBAB (0.4 mol%), CO2 pressure (1 bar), 80 °C, 24 h. | ||||
| 1 | 1′ | TBAB (0.4%) |
|
98 |
| 2 | 1′ | TBAB (0.4%) |
|
95 |
| 3 | 1′ | TBAB (0.4%) |
|
78 |
| 4 | 1′ | TBAB (0.4%) |
|
95 |
| 5 | 1′ | TBAB (0.4%) |
|
90 |
| 6 | 1′ | TBAB (0.4%) |
|
68 |
| 7 | 1′ | TBAB (0.4%) |
|
33 |
Moreover, a detailed comparison between the catalytic activity of 1′ towards the cycloaddition of CO2 to epoxide ring was performed with other reported MOF catalysts in Table S9 (ESI†). A comparative literature survey reveals the superiority of the catalytic performance of 1′ over other reported catalysts in terms of the mild pressure and temperature required for its function.
000 rpm for 5 min. The process was repeated thrice and the recovered catalyst was dried at 80 °C for 6 h. Finally, it was activated under vacuum at 80 °C for 12 h. The activated catalyst was evaluated for the next catalytic cycle under a similar optimized condition with epichlorohydrin. With each cycle, a slight decrease in catalytic activity was observed (Fig. S42, ESI†). Up to the 4th cycle of recyclability, 1′ had efficiently converted 82% (Fig. S37, ESI†) of epichlorohydrin to its corresponding cyclic carbonate. Furthermore, PXRD (Fig. S43, ESI†), ATR-IR (Fig. S44, ESI†), and FE-SEM (Fig. S45, ESI†), analysis were performed on the recycled 1′, and the similarity in results between fresh 1′ and recycled 1′ was assessed. Moreover, the obtained BET surface area of reused 1′ (598 m2 g−1) (Fig. S46, ESI†) and fresh 1′ are quite comparable. Furthermore, ICP-MS analysis was performed on the filtrate obtained after catalytic cycloaddition of CO2 with epichlorohydrin. The results (Table S7, ESI† entry no. 1) revealed that the Zr(IV) ion concentration in the filtrate was insignificant, which nullified the possibility of the Zr(IV) ion leaching from the MOF during the catalytic cycle. These results suggest that structural integrity, crystallinity, and functionality were retained during the catalytic cycles.
![]() | ||
| Scheme 1 A plausible mechanism of fluorine functionalized porous 1′ MOF-mediated catalytic cycloaddition of epoxide and CO2. | ||
By taking the experimental outcomes and former literature reports into account,42,43 we described a plausible mechanism for the cycloaddition in Scheme 1. In the first step, the epoxide will be diffused through the pores of 1′ and undergo weak reversible interaction with the Lewis acidic Zr(IV) sites of the octahedral oxo cluster, followed by adsorption and activation of CO2 molecule in the near vicinity of the catalytic site by Lewis basic –F functionality. In the next step, the ring opening of the epoxide takes place via a nucleophilic attack by the Br− ion of the TBAB co-catalyst on the less hindered site of the epoxide in an SN2 manner. Furthermore, coupling between activated CO2 and the ring-opened epoxide causes the formation of an alkyl carbonate anion. In the last step, an intermolecular ring-closing reaction occurs via a nucleophilic attack by the alkyl carbonate anion on the most electrophilic β-carbon atom, followed by elimination of Br−. In the last step, reductive elimination of five-membered cyclic carbonate occurs to generate 1′ for the subsequent cycle.
| Sl. no. | Hydrocarbon | Catalyst | Conv.b (%) | Selectivityb (%) | Others (%) | |
|---|---|---|---|---|---|---|
| ol | one | |||||
| a Reaction conditions: hydrocarbon (1 mmol), 1′ (10 mg), TBHP (15 μL), acetonitrile (2.5 mL), O2 purged through balloon, 60 °C, 24 h; hydrocarbon (1 mmol), UiO-66-NH2 (10 mg), TBHP (15 μL), acetonitrile (2.5 mL), 60 °C, O2 purged through balloon, 24 h. b Conversion and selectivity were determined by GC–MS. Selectivity to ol/one represents the corresponding alcohol and ketone. | ||||||
| 1 |
|
1′ | 28 | 35 | 60 | 5 |
| 2 | UiO-66-NH2 | 15 | 23 | 63 | 14 | |
| 3 |
|
1′ | 10 | — | 100 | — |
| 4 | UiO-66-NH2 | 2 | — | 100 | — | |
| 5 |
|
1′ | 14 | — | 100 | — |
| 6 | UiO-66-NH2 | 6 | — | 93 | 7 | |
| 7 |
|
1′ | 17 | — | 100 | — |
| 8 | UiO-66-NH2 | 9 | — | 100 | — | |
| 9 |
|
1′ | 69 | 24 | 67 | 9 |
| 10 | UiO-66-NH2 | 12 | 18 | 75 | 7 | |
Considering the observations on the aerobic oxidation of hydrocarbons using 1′, a reaction mechanism is proposed as shown in Scheme S2 (ESI†). The conversion of hydrocarbon (indane) was almost negligible using 1′ as a catalyst without TBHP, while the conversion of indane was appreciable with TBHP. Thus, the Lewis acidic sites (Zr(IV)) in 1′ activate TBHP to afford t-BuOO˙ radicals,44,45 which subsequently abstracts hydrogen from indane to provide carbon radical (indanyl radical). This further reacts with molecular oxygen to give indanylperoxy radical, followed by hydrogen abstraction to provide indanehydroperoxide. Finally, this intermediate affords the expected ol/one mixture with the assistance of Zr(IV) in 1′.
Table S10 (ESI†) compares the activity of 1′ with other hydrophobic MOFs reported in the literature for the oxidation of hydrocarbons. Although this work tested the reactivity of five different hydrocarbons in the presence of 1′, the activity of 1′ may not be compared with previous catalytic data as the reaction conditions, such as the substrate, reaction temperature, and oxidants, are different. In addition, the number of studies on the use of hydrophobic MOFs in the aerobic oxidation of hydrocarbons is limited. In any case, the present work clearly illustrates the importance of hydrophobicity in achieving higher activity compared to non-hydrophobic MOFs. Furthermore, this observation is in good agreement with previous reports.28,29
![]() | ||
| Fig. 5 (a) Changes in fluorescence intensity of 1′ as a function of time. (b) Time-dependent fluorescence saturation plot (the standard deviations of three individual measurements are displayed as the red error bars). The luminescence colour change of 1′ under UV-lamp after injection of ACF is displayed in the inset of Fig. 5(b). | ||
To verify the potential applicability of this sensor, the selectivity of 1′ to detect ACF was evaluated in the presence of various metal ions and herbicides that are commonly present in real samples. The results demonstrated that 1′ exhibited high selectivity for ACF sensing (Fig. 6(a) and Fig. S64–S74, ESI†). However, the addition of 300 μL of ACF solution in the presence of other competitive analytes resulted in similar quenching of fluorescence of the MOF (Fig. 6(b)). Therefore, 1′ demonstrates a strong preference for detecting ACF, even in the existence of its competitive congeners (Fig. S64–S74, ESI†). The error bars in the bar plots depict standard deviation values of three individual measurements which indicate the reproducibility of these measurements. The accuracy of sensing in the intra-day and inter-day (on three distinct days), precision (Table S6, ESI†), and the error of all the measurements (Table S6, ESI†) were assessed to validate the originality and repeatability of the measurements.
The determination of the Stern–Volmer (S–V) quenching constant (Ksv) is crucial to understanding the sensing mechanism responsible for the quenching of fluorescence in the presence of ACF. The determined Ksv also provides valuable information on the sensitivity of the probe for targeted analytes. Here, the Ksv value was computed using the equation: I0/I = Ksv[Q] + 1, where I0 and I represent the emission intensity of 1′ in the presence and absence of the ACF, respectively. [Q] signifies the concentration of the analyte. For ACF, the calculated Ksv value is remarkably high (5.38 × 106 M−1) (Fig. S76 and S77, ESI†). This high Ksv value indicates a strong affinity and sensitivity of 1′ for ACF.
The assessment of a sensor's performance heavily depends on its LOD value. Therefore, calculating the LOD of a sensor is very important. To determine the LOD of 1′ for the sensing of ACF, initially, the standard deviation (σ) of luminescence intensities emitted by 1′ in the absence of any analyte was calculated. Subsequently, a series of fluorescence titration experiments were conducted by progressively diluting the ACF solution. Plotting the concentrations of ACF against the resulting fluorescence emission intensities yielded a linear relationship with a negative slope (k) (Fig. S78, ESI†). Finally, the LOD value was computed using the formula, 3σ/k. The computed LOD value of the MOF to sense the ACF is 1.1 ± 0.2 nM. This LOD value is significantly lower than all the other sensors of ACF (Table S8, ESI†). Additionally, the limit of quantification for this measurement was 3.8 ± 0.6 nM. The average regression equation for the LOD curve was calculated as 15871.6x + 427206.7 with an outstanding average regression coefficient (R2) of 0.996 (Table S4, ESI†). All the statistical results supported the excellent sensitivity of the probe to sense ACF.
A reliable, cost-effective sensor must maintain its quenching effectiveness even after undergoing multiple operational cycles. In this context, we conducted the recyclability test of 1′ for the sensing of ACF up to five cycles. After each cycle of sensing, the MOF material was thoroughly washed with methanol and activated. The quenching efficiency of the probe remained virtually unchanged even after the fifth cycle of sensing (Fig. S79, ESI†). This experiment conclusively demonstrated that 1′ can be effectively recycled to sense ACF. The fluorescence lifetime measurements of the sensor before and after the treatment of ACF were also executed. The obtained lifetime of the MOF before and after the sensing of ACF were 4.6 and 2.4 ns, respectively (Fig. S80 and Table S2, ESI†).
The details of the preparation procedure of the composite are discussed in the ESI† of this manuscript. After synthesis, the successful immobilization of 1′ on the fabric surface was confirmed by PXRD, IR, and FE-SEM analysis of the cotton fabric (Fig. S84–S86, ESI†). Next, various concentrations of ACF were injected into the composites, and after drying, they were placed inside a UV chamber. It was observed that the fluorescence intensity of the composites was quenched to achieve up to the nanomolar concentrations of ACF (Fig. 7). These observations ensure that the composite is capable of detecting the presence of ACF at nanomolar concentrations. Additionally, this robust composite can be reused up to three times without a significant loss in its detection efficiency (Fig. S87, ESI†).
![]() | ||
| Fig. 7 Digital images of 1′@starch@cotton composites under UV-lamp after treatment with different concentrations of ACF solutions. | ||
Numerous phenomena have been documented in the literature, including ground-state complexation (GSC), photoinduced electron transfer (PET), the inner-filter effect (IFE), and Förster resonance energy transfer (FRET), all of which could be accounted for the diminution of fluorescence of 1′ after the addition of ACF in the sensing medium. Based on the change in the lifetime of the excited state, all the above-mentioned phenomena can be categorized into two groups: dynamic and static change in fluorescence. If the fluorophore's lifetime changes after the introduction of an external analyte, the process is considered dynamic, whereas if it remains unchanged, the process is deemed to be static. In the case of IFE and GSC processes, fluorescence lifetime generally remains constant. However, for PET and FRET processes, alterations in lifetime are typically observed.
In the present case, changes in the lifetime of the MOF after the addition of ACF were observed. The lifetime of the MOF dropped significantly from 4.6 ns to 2.4 ns after the addition of ACF in the sensing medium (Fig. S80 and Table S1, ESI†). Such decrease in lifetime values cancelled the possibility of IFE and GSC.
We further investigated the possibility of GSC between the MOF and ACF by examining the solid-state ATR-IR and solid-state UV-Vis spectra of 1′ before and after its treatment with ACF (Fig. S92 and S89, ESI†). Significant changes in the solid-state UV-Vis and IR spectra before and after treatment would be expected for GSC. However, the negligible changes in the solid-state UV-Vis and IR spectra of the analyte-treated 1′ dismissed the possibility of GSC in the sensing medium after the addition of ACF.
The interference of IFE in the fluorescence of a fluorophore can be verified by the shifting of the excitation wavelength of the fluorophore. If IFE is responsible for the quenching of fluorescence of the fluorophore, the quenching efficiency must be decreased after shifting the excitation wavelength of the fluorophore to a higher wavelength region. In the present case, the quenching efficiency remains similar even after shifting the excitation wavelength of the MOF to 390 nm from 350 nm (Fig. S93, ESI†). This experiment provided evidence that IFE is not responsible for the quenching of fluorescence of 1′ in the presence of ACF.
Here, the quenching occurs in a dynamic way which opens the possibility of both FRET and PET mechanisms. PET is a process involving the transfer of excited state electrons, in which the excited state electron of one molecule transfers to the low-lying vacant orbital of another molecule. Here, the quenching of fluorescence of the MOF occurred in the presence of ACF. Therefore, for a feasible PET process, electron transfer should occur from the lowest unoccupied molecular orbital (LUMO) of MOF to the LUMO of ACF. The theoretical location of the highest occupied molecular orbital (HOMO) and LUMO of the linker of the MOF and ACF indicate that the location of the LUMO of ACF is much higher as compared to that of the MOF (function B3LYP and Pople diffuse basis set 6-31G++(d, p) were utilized for all the calculations) (Fig. S94, ESI†). Therefore, the transfer of electrons from the LUMO of MOF to the LUMO of ACF is not energetically favourable. Therefore, PET is not the actual reason for quenching of fluorescence of the MOF in the existence of ACF.
To elucidate the possibility of FRET, we acquired UV-Vis absorption spectra for all tested analytes. Notably, the absorption spectrum of ACF displayed significant overlap with the emission spectrum of the MOF (Fig. S95, ESI†). This observation suggests that FRET could be a probable mechanism for energy transfer from the electron-rich MOF to the electron-deficient ACF; the change in the lifetime values also supports this hypothesis. The red shifting of the emission maxima of 1′ after the addition of ACF indicates the formation of a short-lived weakly interactive species by the close association of ACF and the MOF (required for the FRET process). The formation of such a species in the sensing medium by the close association between the MOF and ACF is responsible for the shift of emission maxima of 1′ by 35 nm.47 Moreover, no other tested analyte's absorption spectrum displays any significant overlap with the emission maximum of the MOF. This could be the reason for the tremendous selectivity in ACF sensing by 1′ against the other competing analytes.
Footnotes |
| † Electronic supplementary information (ESI) available: NMR spectra, PXRD patterns, IR spectra, EDX spectra, TGA curves, FE-SEM images, N2 and CO2 sorption isotherms, catalytic results, sensing plots, fluorescence lifetime plots, recyclability, UV-vis spectrum, comparison tables and statistical results. See DOI: https://doi.org/10.1039/d3tc04140a |
| ‡ Equal contribution. |
| This journal is © The Royal Society of Chemistry 2024 |