Zhefei Guo‡
a,
Geneviève W. Tremblay‡
a,
Jingdan Chen
b and
Shira Joudan
*a
aDepartment of Chemistry, University of Alberta, Edmonton T6G 2G2, Alberta, Canada. E-mail: joudan@ualberta.ca
bDepartment of Chemistry, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA
First published on 6th May 2025
Trifluoromethylphenols (TFMPs) are environmental contaminants that exist as transformation products of aryl-CF3 pharmaceuticals and agrochemicals. Their –CF3 moiety raises concerns as it may form problematic fluorinated transformation products such as the persistent pollutant trifluoroacetic acid (TFA). This study investigates the hydrolysis and spontaneous defluorination mechanisms of 2-TFMP, 3-TFMP, 4-TFMP, and 2-Cl-4-TFMP under environmentally relevant aqueous conditions, and under alkaline pH to investigate the mechanism of defluorination. 3-TFMP did not undergo hydrolysis. The other TFMPs reacted to primarily form the corresponding hydroxybenzoic acids and fluoride. High-resolution mass spectrometry identified a benzoyl fluoride intermediate in the hydrolysis of 4-TFMP and other dimer-like transformation products of the 4- and 2-Cl-4-TFMP. Density functional theory calculations revealed that the key defluorination step likely proceeds via an E1cb mechanism, driven by β-elimination. Experimental and computational results demonstrated substituent-dependent differences in reactivity, and the importance of the deprotonation of TFMPs for the hydrolysis reaction to proceed. These findings provide mechanistic insights into the complete defluorination of TFMPs and broader implications for the environmental defluorination of other PFAS.
Environmental significanceThe environmental transformation of some aryl-CF3 pharmaceuticals and agrochemicals produces trifluoromethylphenols (TFMPs), which are environmental contaminants themselves and can be precursors to the very persistent pollutant trifluoroacetic acid (TFA). Alternatively, this study demonstrates that certain TFMPs can undergo complete hydrolytic defluorination in aqueous solutions, depending on their chemical structure. Understanding what drives this spontaneous defluorination can provide insight into why certain C–F bonds can break, despite the general stability of organofluorine compounds. Designing aryl-CF3 compounds to favor defluorination pathways without TFA formation could harness the benefits of the –CF3 moiety in pharmaceuticals and agrochemicals while reducing their environmental persistence. |
Hydrolysis is an important abiotic transformation pathway for organic pollutants. Previous research observed that 2- and 4-TFMP can undergo hydrolysis resulting in spontaneous defluorination under alkaline conditions.9,13,19–24 Reuben G. Jones first reported the base-mediated defluorination of 4-TFMP in 1947, observing that higher concentrations of 4-TFMP spontaneously defluorinated and polymerized in NaOH solutions, eventually stabilizing with no further defluorination.19 Later, in 1973, Sakai and Santi observed complete hydrolytic defluorination of 2- and 4-TFMP in buffered solutions (pH 6.5–13), yielding 2- and 4-hydroxybenzoic acid (2- and 4-HBA), respectively.20 They proposed that the defluorination mechanism is likely facilitated by β-elimination via a conjugated structure that stabilizes the phenolate anion, as illustrated in Fig. 1C. More recently, environmental chemists, including the Mabury group,13,21,22 Manfrin et al.,23 and Bhat et al.,9 reported similar defluorination reactions. 4-TFMP is the most well-studied TFMP, however, reports on its hydrolytic stability are conflicting. Some suggest 4-TFMP rapidly loses HF upon dissolution in water,13,21 while others propose that spontaneous defluorination requires additional bases that significantly deprotonate the phenol-OH group; otherwise, degradation is minimal, or the hydrolysis defluorination rate is inferred to be extremely slow.20,22
The spontaneous defluorination of certain TFMPs under environmentally relevant conditions contradicts the common perception that C–F bonds are difficult to activate and that fluoride (F−) is a poor leaving group.25 The mechanism of this spontaneous defluorination remains largely unexplored due to the lack of identified intermediates and transformation products. No plausible mechanism has been proposed to explain the reactivity differences caused by substituent effects.
Density functional theory (DFT) calculations are used in environmental chemistry to calculate the thermodynamics and kinetics of reactions, particularly for environmental chemistry reactions involving reactive transient intermediates that cannot be characterized by common analytical methods.26–28 By optimizing reactants, intermediates, products, transition states (TS) and calculating the proposed reaction pathways, it is possible to assess the plausibility of these pathways from both thermodynamic and kinetic perspectives. Combined with experimental data, DFT calculations have been successfully applied to explain the degradation of PFAS26–28 and the hydrolytic defluorination of fluorinated pharmaceuticals.29
This study systematically investigated the mechanism of TFMPs that undergo spontaneous defluorination in aqueous solutions under environmentally relevant conditions to build off previously conflicting reports of TFMP hydrolysis rates.13,21,22 To explore the substituent effects, 2-TFMP, 3-TFMP, 4-TFMP, and 2-Cl-4-TFMP were studied in aqueous solutions (Fig. 1B). pH buffer solutions from pH 6.2 to 10.8 were used to study the effect of deprotonation on hydrolysis kinetics, and Orbitrap high-resolution mass spectrometry (HRMS) was used to identify transformation products. The most plausible spontaneous defluorination mechanism was derived through DFT calculations combined with experimental data.
At set time points for each experiment, 1 mL aliquots were mixed with a volume of 1 M HCl to quench the experiments prior to analysis: 5 μL for pH 6.2–7 experiments, 10 μL for pH 7.9 experiments, and 15 μL for pH 10–10.8 experiments. The concentrations of TFMPs and their corresponding hydrolysis products hydroxybenzoic acids (HBAs) were monitored by high performance liquid chromatography (HPLC) coupled with a UV detector. Detailed instrumental conditions are provided in the Table S1.†
Plotting ln(Ct/C0) versus time (t) provided the pseudo-first-order rate constants, k (eqn (1)). The half-lives (t1/2) of the compounds were determined from eqn (2). Eqn (3) describes the fraction of deprotonated TFMPs (in phenolate form, fphenolate) as a function of aqueous solution pH based on Henderson–Hasselbalch equation. Eqn (4) explains the degradation of TFMPs by using the fraction of TFMPs that dissociate into phenolates, where kphenolate is the rate constant when TFMPs are 100% deprotonated in their phenolate form. The Eyring equation (eqn (5)) was used to plot ln(k/T) with respect to 1/RT for each experiment (eqn (6)), where k is the pseudo-first order rate constant (s−1); T is the absolute temperature in K; R (8.314 J (mol−1 K−1)) is the gas constant; kB (1.38 × 10−23 J K−1) is the Boltzmann constant; h (6.626 × 10−34 J s) is Planck's constant; κ is the transmission coefficient (close to 1 in most cases). The slope of the eqn (6) fitting plot provided the experimental free energy of activation (ΔG‡, kcal mol−1).
![]() | (1) |
![]() | (2) |
![]() | (3) |
k = kphenolate × fphenolate | (4) |
![]() | (5) |
![]() | (6) |
To determine the fluorine mass balance of pH 10 hydrolysis experiments, F− formation was tracked by the Fisherbrand Accumet AB250 pH/ISE meter and the Accumet combination fluoride ion selective electrode. Fluorine mass balance experiments were run in duplicates, and 1 mL aliquots were combined with 1 mL of the total ionic strength adjustment buffer (Fisher Scientific) prior to analysis. F− calibration standards were prepared from NaF dissolved in ultrapure water. TFMPs were also quantified by HPLC-UV in these separate experiments.
To determine how pH may affect transformation products formation, aliquots taken at different time points during pH 6.2 and pH 10.8 experiments were diluted 10 times in 50:
50 (v/v) methanol
:
H2O and analyzed by ultra-high performance liquid chromatography (UHPLC) coupled with an Orbitrap HRMS. For pH 10 experiments, samples were only taken at the beginning and end. Detailed UHPLC-Orbitrap-HRMS parameters are listed in the Table S2.† Suspected transformation products are listed in Table S4.†
In situ 19F NMR was also conducted in an attempt to detect highly reactive fluorinated intermediates. Instrumental details are described in ESI-3.†
We specifically considered a hybrid solvation model for key elementary reactions to account for solvation with water. For the explicit solvation, one, two, five, or six explicit H2O molecules were added to calculations. For the implicit solvation, the solvation free energy was taken into account in the Gibbs free energy (Gf) of a molecule by using solvation model density (SMD) model34 as indicated by Chen et al.29 and Qin et al.35 Gf values were also corrected for the standard state in solution as indicated by Zhang et al.26,27 For reactions with explicit H2O molecules, the Gf of H2O was corrected for its concentration (55.3 M) deviation from its standard state (1 M).29 Refer to ESI-4 and 11† for more details.
![]() | ||
Fig. 2 Hydrolysis half-lives and rate constants of 4-TFMP, 2-TFMP, and 2-Cl-4-TFMP based on buffer pH. The solid lines for three TFMPs are fitting the curve of eqn (3) and (4). Note that the hydrolysis half-life time is plotted on the left y-axis while the hydrolysis rate constant is plotted on the right y-axis. Both y-axes are in logarithmic scale. Results are expressed as mean ± standard deviation; error bars are within the points. |
Fig. 2 displays the aqueous hydrolysis half-life times and rate constants of 2-TFMP, 4-TFMP and 2-Cl-4-TFMP as a function of pH. The good fitting results (R2 > 0.95) of eqn (3) and (4) to the experimental data indicates that the aqueous hydrolysis of 2-TFMP, 4-TFMP, and 2-Cl-4-TFMP is almost entirely driven by the deprotonated species, the phenolate. When fully deprotonated, each TFMP has its own fundamental hydrolysis rate constant, kphenolate, with k4-TFMP-phenolate (0.62 ± 0.05) h−1 > k2-Cl-4-TFMP-phenolate (0.115 ± 0.004) h−1 > k2-TFMP-phenolate (0.0671 ± 0.0007) h−1 (Table 1). The aqueous hydrolysis reaction rates of these three TFMPs are determined by a combination of the fraction of deprotonation and the fundamental hydrolysis rate constants of the phenolates. The fitting curves of 4-TFMP and 2-Cl-4-TFMP intersect at ∼pH 8, which implies prior to pH 8, the hydrolysis reaction rate of 2-Cl-4-TFMP is faster than that of 4-TFMP. The acid-base dissociation equilibrium of these three TFMPs dictates that a certain amount of phenolate always exists, and thus its hydrolysis can thermodynamically push the acid–base equilibrium forward. However, not all reactions are favorable kinetically. For example, in the pH 6.2 buffer over five days, only 0.5% of 4-TFMP exists in its phenolate form and slow hydrolysis was observed (2.1 × 10−3 ± 0.05 × 10−3 h−1), while 1.2% of 2-TFMP exists in its phenolate form, but no hydrolysis was observed. This is because the fundamental hydrolysis rate of the fully deprotonated 2-TFMP is too slow (nine times slower than 4-TFMP).
Compound | pKa | Buffer/pH/temperature | Rate constant (h−1) | Half-life (h) |
---|---|---|---|---|
N.A. = not available; acalculated from b; bfrom ref. 22; cfrom ref. 20; values without superscripts are experimental results obtained from triplicates in this work; dcalculated from our experimental results across a range of 5 pH. | ||||
![]() |
8.51 | 10 mM phosphate/7/22 °C | 0.0147 ± 0.0005 | 47.0 ± 1.5 |
10 mM carbonate/10/22 °C | 0.612 ± 0.008 | 1.13 ± 0.01 | ||
10 mM carbonate/10/30 °C | 1.98 ± 0.03 | 0.350 ± 0.005 | ||
10 mM carbonate/10.8/22 °C | 0.709 ± 0.010 | 0.976 ± 0.014 | ||
50 mM borate/8.5/N.A. | a0.66 | b1.1 | ||
200 mM borate/10/30 °C | c2.3 | c0.30 | ||
200 mM phosphate/7/30 °C | c0.073 | c9.5 | ||
k4-TFMP-phenolated | 0.621 ± 0.047 | 1.11 ± 0.08 | ||
![]() |
8.12 | 10 mM phosphate/7/22 °C | 0.0051 ± 0.0002 | 137 ± 4 |
10 mM carbonate/10/22 °C | 0.0661 ± 0.0008 | 10.5 ± 0.1 | ||
10 mM carbonate/10/30 °C | 0.217 ± 0.001 | 3.19 ± 0.01 | ||
10 mM carbonate/10.8/22 °C | 0.0674 ± 0.0012 | 10.3 ± 0.2 | ||
200 mM borate/10/30 °C | c0.26 | c2.7 | ||
200 mM phosphate/7/30 °C | c0.0093 | c74 | ||
k2-TFMP-phenolated | 0.0671 ± 0.0007 | 10.3 ± 0.1 | ||
![]() |
7.09 | 10 mM phosphate/7/22 °C | 0.0694 ± 0.0006 | 9.99 ± 0.09 |
10 mM carbonate/10/22 °C | 0.109 ± 0.004 | 6.36 ± 0.24 | ||
10 mM carbonate/10.8/22 °C | 0.121 ± 0.002 | 5.72 ± 0.09 | ||
10 mM carbonate/10/30 °C | 0.431 ± 0.004 | 1.61 ± 0.01 | ||
k2-Cl-4-TFMP-phenolated | 0.115 ± 0.004 | 6.02 ± 0.21 |
When tested at an elevated temperature of 30 °C in pH 10 solution, we were able to compare our 4-TFMP and 2-TFMP degradation rates with those reported by Sakai and Santi more than 50 years ago.20 We found that our measured rates closely align with theirs (Table 1), despite the different buffers used: their experiments in 200 mM borate buffer, and ours in 10 mM carbonate. To the best of our knowledge, there is no previous literature to compare to our 2-Cl-4-TFMP results.
![]() | ||
Fig. 3 Simplified proposed reaction mechanism for 4-TFMP, 2-TFMP, and 2-Cl-4-TFMP spontaneous aqueous defluorination. A detailed arrow-pushing mechanism can be found in Fig. S7.† |
UHPLC-Orbitrap-HRMS and in situ 19F NMR were used to search for additional transformation products, either as intermediates of the reaction from TFMPs to HBAs, or other stable products of the reaction. The benzoyl fluoride TP-4-1 was only observed in HRMS (Fig. 3) in the beginning of 4-TFMP hydrolysis experiments at pH 10. The isotope pattern of the benzoyl fluoride aligns with its molecular formula C7H4FO2− (Fig. S3†). While benzoyl fluorides have characteristic positive chemical shifts in 19F NMR, in situ 19F NMR did not detect any signal in the positive region (Fig. S1†), likely because its lifetime is shorter than the acquisition time between typical 19F NMR arrays and their steady-state concentrations are very low. The observation of this benzoyl fluoride in the HRMS supports the hypothesized pathway where deprotonated 4-TFMP first eliminates F− and forms a highly reactive conjugated electrophile quinone difluoromethide. This quinone then forms an aryl difluoromethanol (–CF2OH) compound through nucleophilic addition with H2O or OH−, and unstable –CF2OH subsequently eliminates HF to form TP-4-1 (Fig. 3). Analogous benzoyl fluoride transformation products were not observed for other TFMPs, either due to lower formation, or decreased stability of the benzoyl fluorides, resulting in lower concentrations in our samples.
The HRMS of 4-TFMP and 2-Cl-4-TFMP hydrolysis experiments identified novel dimer-like transformation products in addition to HBAs, while for 2-TFMP only 2-HBA was detected (Fig. 3). TP-4-3, TP-4-4, TP-2Cl-2, and TP-2Cl-3 are ester transformation products whose exact mass was identified (Table S4†) and peak areas increased over time during hydrolysis experiments (Fig. 4 and S2†). TP-4-3 was also observed in in situ 19F NMR with an increasing peak over time near the 4-TFMP chemical shift (Fig. S1†). At what is labelled t = 0, certain transformation products, such as HBAs and trifluoromethyl-containing phenolic esters (TP-4-3 and TP-2Cl-2) are present, because it is not possible to measure a true time zero for these experiments due to time required for sampling. However, these products were not detected in fresh aqueous stock solutions prepared in ultrapure water used for the experiments nor in the methanol calibration standards, thus we report them as transformation products.
We propose the esterification in TFMP hydrolysis only occurs through the deprotonated phenolate groups of either TFMP or HBA reacting via a conjugate addition with quinone difluoromethides or via a nucleophilic addition–elimination mechanism with the benzoyl fluoride intermediate (Fig. 3). Results across a range of pH support that the esterification products only form when the TFMP or HBA phenolates are present (Fig. S2 and Table S3†). 2-TFMP does not form an ester analogous to TP-4-4, because the high pKa of phenol-OH in 2-HBA (salicylic acid) makes them fully protonated (see Table S3†). The lack of the ester analogous to TP-4-3 for 2-TFMP may be due to the steric hindrance caused by the ortho-position of the fluorinated moiety relative to the phenol-OH. Our group's recent study on the photochemistry of 4-TFMP and Zhang et al.'s study on the chlorination of parabens both identified similar coupling products.10,36 The formation of these esterification products may be limited in real-world aqueous environments where concentrations are lower, but it is also possible that other aromatic pollutants or organic matter with nucleophilic group (–NH2, –OH) in the environment may react with the reactive intermediates generated during the hydrolysis.
![]() | ||
Fig. 5 DFT pathways with the minimum energy barriers for 4-, 2-Cl-4-, 2-TFMP aqueous spontaneous defluorination reactions rate-determining steps under the hybrid solvation model. |
The DFT-calculated ΔG‡ for the hydrolytic defluorination of 4-TFMP (21.2 kcal mol−1) and 2-TFMP (22.6 kcal mol−1) closely matched the experimental values of 22.6 kcal mol−1 and 23.5 kcal mol−1, respectively (Table 2). The calculated ΔG‡ for 2-Cl-4-TFMP falls within the standard deviation range of its experimentally determined ΔG‡. 4-TFMP is the most reactive when fully deprotonated in our laboratory experiments, and has the lowest ΔG‡ of the three TFMPs determined both experimentally and theoretically. 2-Cl-4-TFMP is more reactive than 2-TFMP in hydrolysis, however, based on our ΔG‡ determination, we cannot attribute the faster reactivity to a lower energy rate-determining step. This is because of the larger relative error of our experimental ΔG‡ of 2-Cl-4-TFMP, compared to the other TFMPs. DFT calculations reported the highest ΔG‡ for 2-Cl-4-TFMP, but this does not align with our experimental observations, which suggests the current hybrid solvation model used may be insufficient to fully capture the real conditions. The rate-determining step transition state structures for three TFMPs shared similar key features: stretched C–F bonds, hydrogen-bond interactions between explicit water molecules and the phenolate, and stabilization of the transition state through interactions between the water O–H and the departing F− (Fig. 5).
TFMP-phenolate | ΔG‡exp | R2 | Minimum ΔG‡DFT scenario n(H2O) | ΔG‡DFT |
---|---|---|---|---|
2-TFMP | 23.5 ± 1.6 | 0.995 | 6 | 22.6 |
4-TFMP | 22.6 ± 1.8 | 0.999 | 5 | 21.2 |
2-Cl-4-TFMP | 22.8 ± 3.9 | 0.971 | 5 | 25.4 |
![]() | ||
Fig. 6 (A) Natural population analysis (NPA) of –CF3's β-carbons and the bond dissociation energy of C–F bonds in TFMPs (deprotonated, phenolates); CYLview was used to visualize the structures.38. (B) Downhill rearomatization of quinone difluoromethide intermediate for 4-TFMP. |
In Fig. 6B, the quinone difluoromethide intermediate IM-4-1 can undergo a barrierless reaction with OH− to form the same IM-4-2 (ΔG = −39.1 kcal mol−1, see potential energy surface scan results in Fig. S11†), a rearomatization process that is both thermodynamically and kinetically favorable. For 2-Cl-4-TFMP and 2-TFMP, the barrierless reaction between OH− and their corresponding quinone difluoromethides is also the more favored pathway (illustrated in Fig. S10,† with potential energy surface scan results shown in Fig. S11†). Combined with expected high OH− concentration in the basic system, these quinone difluoromethides are highly reactive electrophiles, rapidly captured by OH− or other high concentration nucleophiles such as H2O or deprotonated TFMPs substrates, making them undetectable via regular HRMS or 19F NMR.
For the intermediate IM-4-2, the unstable –CF2OH structure readily eliminates HF via a hydrogen-bonding network. Adding explicit water molecules to the TS of this HF-elimination reduced ΔG‡ from 24.8 to 13.3 kcal mol−1, consistent with prior studies.26 Alternatively, IM-4-2 can undergo intramolecular proton transfer to form IM-4-3, where the negatively charged oxygen in –CF2O− facilitates the leaving of F− with ΔG‡ of only 8.2 kcal mol−1, making it highly favorable kinetically. Similar scenarios are observed for 2-Cl-4-TFMP and 2-TFMP. After the removal of first two F−, all three TFMPs generate the corresponding (deprotonated) benzoyl fluorides, which reacts either with H2O/OH− to eliminate the last F− and to form benzoic acids or with other nucleophiles such as phenol-O− to form dimer-like transformation products.
Although one carbon involved in the β-elimination defluorination reaction of TFMPs is an aryl-carbon (sp2 hybridized), and many PFAS molecules are composed of alkyl-carbons (sp3 hybridized), similar β-elimination defluorination mechanisms are not uncommon in PFAS chemistry and usually occur in the early stages of the overall reaction. For instance, perfluorooctanoic acid (PFOA) can undergo decarboxylation in polar aprotic solvents, forming a perfluoroalkyl carbanion that drives β-elimination of an adjacent C–F bond to give a perfluoroalkene.28 DFT calculations revealed a ΔG‡ of 19.5 kcal mol−1 for this β-elimination step,28 comparable to the ΔG‡ values observed for TFMPs in our study. For polyfluorinated compounds, it has been reported that the presence of easily removable H or Cl atoms can indirectly trigger defluorination.39,40 Multiple electron-withdrawing C–F bonds can weaken other chemical bonds, facilitating the dissociation. In this work, the phenol-O–H bonds in TFMPs are more readily dissociated than the equivalent non-fluorinated methylphenols (cresols). For example, the pKa of 4-TFMP is 8.51 compared to 10.25 for 4-methylphenol, due to the trifluoromethyl group stabilizing the carbanion. For example, 6:2, 8:2, 10:2 fluorotelomer carboxylic acid (n:2 FTCA, Cn−1F2n−1CF2CH2COOH, n = 6/8/10) can transform into n:2 fluorotelomer unsaturated carboxylic acid (n:2 FTUCA, Cn−1F2n−1CFCHCOOH, n = 6/8/10) biotically or abiotically.41,42 This abiotic transformation is facilitated by a base,41,42 likely through the deprotonation of a C–H adjacent to the carboxyl group, leading to β-elimination and the formation of an α,β-unsaturated carbonyl compound. Similarly, 8:2 fluorotelomer aldehyde (8:2 FTAL, C7F15CF2CH2CHO) is unstable in water, and loses HF to produce the 8:2 fluorotelomer α,β-unsaturated aldehyde (8:2 FTUAL, C7F15CF
CHCHO).43 A recent study suggest that an unsaturated perfluorinated compound E-perfluoro-4-methylpent-2-enoic acid [(CF3)2CFCF
CFCOOH] can undergo microbial defluorination, where similar base-facilitated β-elimination defluorination contributes as part of the mechanism.39
HBAs and F− were the major products for the three TFMPs, as confirmed by their molar yields and fluorine mass balance. UHPLC-Orbitrap-HRMS and in situ 19F NMR identified new transformation products, including a benzoyl fluoride intermediates for 4-TFMP and dimer-like products for 4-TFMP and 2-Cl-4-TFMP, offering additional insights into the reaction mechanisms.
The key rate-determining step for defluorination involved β-elimination, likely via an E1cb mechanism, initiated by the deprotonated phenolate. DFT calculations provided detailed insights into this process, highlighting the critical role of water molecule solvation in stabilizing the phenolate and facilitating β-elimination. The reactivity was influenced by both the C–F bond BDE and the negative charge distribution on the phenolate of the TFMPs. Additionally, DFT calculations revealed that following the rate-determining step, the rearomatization of the reactive quinone difluoromethide intermediate was highly favorable both thermodynamically and kinetically. These findings enhance our understanding of defluorination pathways for TFMPs and provide insights into the defluorination chemistry of other PFAS. This may be used to design aryl-CF3 compounds that favour defluorination in the environment, thus minimizing their environmental impact. As aryl-CF3 contaminants are more and more incorporated in the environmental analysis, our research findings also indicated that if TFMPs are to be targeted as analytes, it is important to store samples at low pH and avoid using base during sample preparation to minimize loss via hydrolysis.
Footnotes |
† Electronic supplementary information (ESI) available: Ref. 44–54. See DOI: https://doi.org/10.1039/d4em00739e |
‡ Z. G. and G. W. T. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |