Rana Dalapati,
Jiangfan Shi,
Matthew Hunter and
Ling Zang
*
Department of Materials Science and Engineering, University of Utah, Salt Lake City, Utah 84112, USA. E-mail: lzang@eng.utah.edu
First published on 15th July 2025
Perfluorooctanoic acid (PFOA), a synthetic compound belonging to the per- and polyfluoroalkyl substances (PFAS) family, is notorious for its environmental persistence, bioaccumulation potential, and adverse health effects, posing a major challenge to environmental safety. Effective removal and detection of PFOA remains a challenge for conventional capture materials due to its unique molecular structure. In this study, we present a dual-functional metal–organic framework (MOF), UiO-66-N(CH3)3+, specifically designed for selective removal and detection of PFOA in water. This MOF is synthesized through the post-synthetic modification of UiO-66-NH2(Zr) with methyl iodide, introducing partially quaternized ammonium groups that enable ion-exchange functionality. The cationic ammonium groups significantly enhance electrostatic interaction with the anionic PFOA, leading to improved affinity and selectivity. As a result, UiO-66-N(CH3)3+ exhibits outstanding adsorption performance, achieving a high adsorption capacity of 1178 mg g−1 as estimated using the Langmuir isotherm model, along with over 99% removal efficiency within 5 minutes from a 50 ppb PFOA solution. Beyond its high sorption capability, the same MOF is also developed into a highly efficient fluorescence “turn-on” sensor for PFOA detection via a straightforward indicator displacement assay (IDA). In this approach, the MOF is initially loaded with the anionic dye sulforhodamine B (SRB), replacing the original iodide counterions. When bound within the MOF, SRB is nonfluorescent; however, upon exposure to PFOA, it is displaced and regains its strong fluorescence in solution. This rapid fluorescence “turn-on” response enables effective detection of PFOA with both high sensitivity and selectivity. The dual-functional MOF system described herein offers a promising strategy for the integrated detection and removal of PFOA from water, providing a simple yet powerful framework for designing multifunctional MOF-based adsorbents and sensors for PFAS pollutants.
Among various remediation strategies, adsorption is considered practical and environmentally sustainable. Commercial sorbents like activated carbon (AC)7 and anion exchange resins (AERs)8 show good PFAS uptake but are limited by slow kinetics,9,10 low reusability,11,12 or poor selectivity.13 To overcome these limitations, synthetic sorbents with functional diversity and tailored architectures have been developed to promote specific PFAS interactions.14,15 However, the mechanisms underlying selectivity remain unclear, and rational design of PFAS-specific sorbents remains an ongoing challenge.
Metal–organic frameworks (MOFs), a class of porous crystalline materials built from metal nodes and organic linkers, have shown strong potential for PFAS removal.16–18 Their high surface area, large internal voids, and tunable chemistry allow for efficient PFAS encapsulation and adsorption via multimodal interactions.19,20 These unique attributes make MOFs highly promising candidates for developing selective and versatile PFAS adsorbents suitable for diverse environmental conditions. Moreover, post-synthetic modification (PSM) offers a powerful strategy for introducing various functional groups into MOFs to enhance their PFAS uptake capacity.21,22 Recent studies on various non-MOF sorbent materials have demonstrated that surface modification with cationic functional groups enhances PFAS capture through strong electrostatic interactions.23–25 Consequently, the design of MOFs incorporating cationic side groups through PSM presents a straightforward and effective approach for enhancing PFAS adsorption performance.
Building on this concept, we developed a dual-functional MOF platform for both PFOA adsorption and detection. Through post-synthetic modification, cationic ammonium groups were introduced into MOF UiO-66-NH2, to create UiO-66-N(CH3)3+ with enhanced ion exchange capacity (Scheme 1). This modified MOF exhibited significantly higher PFOA uptake compared to the unmodified MOF, activated carbon, and other conventional sorbent materials. An impressive adsorption capacity of 1178 mg g−1 was achieved, as estimated by fitting the experimental data to the Langmuir isotherm model, along with fast adsorption kinetics, excellent reusability, and strong salt tolerance.
![]() | ||
Scheme 1 Schematic illustration of the synthesis of UiO-66-N(CH3)3+, adsorption of PFOA via ion exchange, and IDA-based fluorescent sensor for PFOA. |
Beyond PFAS removal, the intrinsically high adsorption capacity of cationic MOFs can also be harnessed for developing highly sensitive PFAS sensors. Accurate quantification of PFAS, particularly PFOA, is critical for assessing environmental contamination, understanding exposure pathways, and evaluating health risks. However, conventional detection methods, such as liquid chromatography coupled with mass spectrometry, are limited by complex sample preparation and lengthy analysis times, rendering them impractical for real-time, on-site monitoring.26,27 These challenges have spurred the development of alternative strategies, with fluorescence-based sensors emerging as a promising solution due to their high sensitivity, rapid response, and cost-effectiveness.28–33
MOF-based sensors offer dual functionality by combining PFAS capture with detection. Their high porosity, tunable surface chemistry, and inherent fluorescence responsiveness enable selective binding and real-time monitoring of PFAS via signal modulation. A particularly innovative strategy for designing MOF-based sensors involves the use of indicator displacement assays (IDA), though it remains underexplored for PFAS detection.34–36 In this work, we developed an IDA-based fluorescence sensing strategy using sulforhodamine B (SRB) as an anionic indicator, which is released upon PFOA binding, generating a distinct fluorescence “turn-on” signal (Scheme 1). To the best of our knowledge, this represents the first example of a MOF-based platform that integrates high-capacity PFOA adsorption with IDA-based fluorescence sensing, offering a simple, effective, and dual-functional solution for addressing PFAS contamination.
Quantitative analysis of PFOA at high concentrations (in the ppm range) was performed using 19F NMR. Briefly, 700 μL of the filtered supernatant was transferred to an NMR tube, followed by the addition of 50 μL of trifluoroacetic acid (TFA) in deuterium oxide (D2O) as an internal standard. The 19F NMR measurements were conducted with a 15-second relaxation delay to ensure accurate signal integration.40 The concentration of PFOA before and after adsorption was determined using eqn (1):41
[PFOA] = IPFOA/ITFA × NTFA/NPFOA[TFA] | (1) |
For trace-level quantification (50 ppb), PFOA concentrations were determined using LC-MS, for which a calibration curve was established using PFOA standards ranging from 0 to 80 ppb (Fig. S10b, ESI†).
The adsorption kinetic data were fitted using the pseudo-first-order (PFO) model for low PFOA concentrations (<50 ppb) and the pseudo-second-order (PSO) model for higher concentrations in the ppm range, as described by eqn (2) and (3), respectively:
qt = qe(1 − e(−k1·t)) | (2) |
qt = (k2·qe2·t)/(1 + k2·qe·t) | (3) |
Adsorption isotherm studies were conducted using a fixed MOF dose of 500 mg L−1 and varying initial PFOA concentrations (0, 100, 200, 400, 600, 800, and 1000 ppm). The mixtures were stirred for 60 minutes to reach the adsorption equilibrium, and the adsorption capacities at equilibrium (qe) were fitted to the Langmuir isotherm model:
qe = (Qmax·KL·Ce)/(1 + KL·Ce) | (4) |
PFOA removal efficiency of UiO-66-N(CH3)3+ was calculated according to eqn (5):42
![]() | (5) |
To assess the reusability of UiO-66-N(CH3)3+ for adsorption of PFOA, the MOF was recovered by centrifugation and subsequently treated with 50 mL of a washing solution composed of 0.2 M HCl and methanol (30:
70 v/v). The mixture was sonicated at room temperature for 30 minutes, after which the solid was separated again by centrifugation. This washing process was repeated three times to ensure effective recovery of PFOA through ion exchange with chloride ions. The regenerated UiO-66-N(CH3)3+ was then dried under vacuum at 60 °C for 2 hours before being reused in subsequent adsorption cycles.
UiO-66-NH2(Zr) was selected as the starting material due to its well-established chemical stability and the reactivity of its amine functional groups, which enable diverse post-synthetic transformations.43–45 The PXRD pattern of UiO-66-N(CH3)3+ closely matches that of the parent UiO-66-NH2(Zr), with minimal loss of crystallinity, indicating that the MOF retaines its structural integrity throughout the quaternization process (Fig. 1a). Furthermore, the PXRD pattern of UiO-66-N(CH3)3+@SRB remained largely unchanged, confirming that the anion exchange with SRB did not significantly impact the MOF's crystalline framework.
![]() | ||
Fig. 1 (a) PXRD patterns, (b) FT-IR spectra, and (c) N2 adsorption isotherms as measured for UiO-66-NH2 and UiO-66-N(CH3)3+ (free or loaded with SRB). |
FT-IR analysis was performed to gain insight into the cationic modification of the MOF and the subsequent ion exchange process involving SRB molecules. The FT-IR spectrum of the unmodified UiO-66-NH2(Zr) displays characteristic peaks, including a C–O stretching vibration at 1384 cm−1, an O–Zr–O vibration at 664 cm−1, and prominent bands at 3316 cm−1 and 3480 cm−1 corresponding to the –NH2 functional groups, confirming the successful formation of the MOF framework (Fig. 1b and S2, ESI†).46–48 After quaternization to form UiO-66-N(CH3)3+, a new peak appeared at 1418 cm−1, attributed to –CH3 vibrations. In addition, the C–N stretching vibration, originally observed at 1258 cm−1 in UiO-66-NH2, shifted to 1280 cm−1 in UiO-66-N(CH3)3+, indicating the formation of a new quaternary ammonium group. Importantly, the O–Zr–O vibration at 664 cm−1 remained unchanged, suggesting that the MOF framework remained intact during the quaternization process. The FT-IR spectrum of UiO-66-N(CH3)3+@SRB exhibited new peaks at 1180, 1130, 1075, and 1029 cm−1, which can be attributed to vibrational modes of the SRB molecules. These results confirm the successful incorporation of SRB within the MOF structure (Fig. S3, ESI†).
The Brunauer–Emmett–Teller (BET) surface area was measured via nitrogen sorption experiments to assess changes in porosity of the MOF at different stages: prior to quaternization, after quaternization, and following anion exchange with SRB fluorophores. As shown in Fig. 1c, the MOF retained significant porosity even after the introduction of quaternized amine functional groups. The surface area decreased from 1007 m2 g−1 for UiO-66-NH2(Zr) to 722 m2 g−1 for UiO-66-N(CH3)3+, which is attributed to reduced internal void space resulting from the incorporation of bulky quaternary ammonium groups. A slight further decrease in surface area to 688 m2 g−1 was observed for UiO-66-N(CH3)3+@SRB, compared to the quaternized MOF with iodide as the counterion. This modest reduction is likely attributed to the larger size of SRB relative to iodide.
The extent of the quaternization reaction was confirmed through detailed 1H NMR analysis. For this purpose, the MOF powder was digested using hydrofluoric acid, and the resulting solution was dissolved in DMSO-d6 for NMR measurements. The 1H NMR spectrum of the digested parent UiO-66-NH2(Zr) MOF displayed characteristic peaks at 7.01, 7.36, and 7.75 ppm, corresponding to the three aromatic protons of the aminoterephthalate ligand (Fig. S4, ESI†).49 Following modification, a new peak appeared at 7.19 ppm, consistent with the formation of the quaternized ammonium ligand. Additional peaks at 7.05 and 7.08 ppm were observed, which are indicative of residual primary and secondary amine-containing species, respectively.39 By analyzing the integrated peak areas corresponding to the NH2-BDC ligand and the quaternized products, the conversion efficiency of the quaternization reaction was estimated to be approximately 52%.
The same digestion protocol was employed to verify the successful exchange of SRB within the MOF structure. The 1H NMR spectrum of the digested UiO-66-N(CH3)3+@SRB sample exhibited new peaks at 1.19, 3.62, 6.89, 7.02, 7.14, 7.71, and 8.26 ppm, all of which matched the spectrum of free SRB molecules (Fig. S5, ESI†). These results confirm the successful formation of UiO-66-N(CH3)3+@SRB via ion exchange with the SRB fluorophore.
SEM imaging was used to examine the morphology of both UiO-66-NH2(Zr) and UiO-66-N(CH3)3+. As shown in Fig. S6 (ESI†), both samples exhibited uniform distribution of MOF particles with sizes below 100 nm. The small particle size enables stable and homogeneous dispersion in solution, which is advantageous for sensor testing and facilitates quantitative analysis with high reproducibility. Clearly, as evidenced by the SEM images, the morphology of the MOF is well preserved following post-synthetic modification.
The adsorption isotherm of PFOA on UiO-66-N(CH3)3+ was established using experimental data collected over an initial concentration range of 0 to 1000 ppm. The data were fitted to the Langmuir isotherm model (Fig. 3c), yielding a high correlation coefficient (R2 = 0.99), indicating excellent model agreement. The maximum adsorption capacity for PFOA was calculated to be 1178 mg g−1 from the Langmuir model, demonstrating the outstanding performance of the material. To the best of our knowledge, this remarkably high uptake surpasses that of most previously reported adsorbents, including other MOF-based materials.53,55–58
To assess the performance of UiO-66-N(CH3)3+ in the presence of salts commonly found in real water samples, the same PFOA adsorption experiments were conducted in the presence of various potentially interfering salts (1000 ppm), including KCl, CaCl2, MgCl2, Na2SO4, and Zn(NO3)2. As shown in Fig. S8 (ESI†), UiO-66-N(CH3)3+ maintained a high PFOA adsorption capacity despite the presence of these salts at high concentrations. This result highlights the MOF's exceptional selectivity for PFOA, even in complex aqueous environments with high ionic strength.
The exceptional PFOA adsorption capacity of UiO-66-N(CH3)3+ prompted further evaluation of its removal efficiency. As shown in Fig. 4, the MOF demonstrated an outstanding removal efficiency of over 97% for a PFOA solution with an initial concentration of 100 ppm. To assess its performance under more environmentally relevant and scalable conditions, a lower PFOA concentration of 50 ppb was also tested using a reduced adsorbent dosage of 20 mg L−1. Remarkably, LC-MS analysis confirmed that over 99% of PFOA was removed from water within just 5 minutes (Fig. S9, ESI†), underscoring the material's rapid and highly effective adsorption capability at trace levels. The time-dependent adsorption behavior at low PFOA concentrations followed a pseudo-first-order kinetic model (Fig. S10, ESI†), in contrast to the pseudo-second-order kinetics observed at higher concentrations. This shift likely arises from the change in the adsorption equilibrium. At low PFOA concentrations, the number of available MOF adsorption sites greatly exceeds the number of PFOA molecules, allowing the concentration of adsorption sites to be treated as effectively constant. Under these conditions, the adsorption rate is primarily governed by the PFOA concentration, resulting in pseudo-first-order kinetics. In contrast, at higher PFOA concentrations, the assumption of constant adsorption site concentration no longer holds, and the adsorption behavior is better described by the pseudo-second-order model.
![]() | ||
Fig. 4 19F NMR spectrum of a 100 ppm PFOA solution before (bottom) and after (top) adsorption by UiO-66-N(CH3)3+ (500 mg L−1). |
To assess its practical applicability, the reusability of UiO-66-N(CH3)3+ was also investigated through repeated adsorption–desorption cycles. After each adsorption step, the recovered MOF was washed with a 30:
70 (v/v) mixture of 0.2 M HCl and methanol, then dried for reuse. As illustrated in Fig. S11a (ESI†), the material retained over 93% of its original PFOA adsorption capacity after five consecutive cycles. The slight decrease in efficiency was primarily attributed to minor material losses during washing and centrifugation. Notably, the removal efficiency remained consistently above 97% throughout all cycles, as confirmed by 19F NMR analysis (Fig. S11b, ESI†).
These results underscore the excellent reusability, chemical stability, and structural integrity of UiO-66-N(CH3)3+ for PFOA removal. The material's consistent performance across multiple cycles highlights its robustness and strong potential for real-world water purification and continuous environmental remediation applications.
Further elemental mapping and EDS analysis were conducted to investigate the anion exchange between PFOA and iodide. Fig. S14 and S15 (ESI†) present the elemental mapping and EDS spectra of UiO-66-N(CH3)3+ prior to PFOA uptake, confirming the presence of iodide as the initial counterion. Following treatment with 1000 ppm PFOA, the elemental mapping and EDS spectra (Fig. S16 and S17, ESI†) revealed a strong fluoride signal, indicating substantial adsorption of PFOA via replacement of the original iodide anions. As summarized in Table S1 and Fig. S18 (ESI†), the iodine content within the MOF decreased progressively with increasing PFOA concentration, while the fluorine content simultaneously increased. Consequently, the fluoride-to-iodide (F/I) ratio showed a consistent upward trend with higher PFOA loading, providing further evidence for an anion exchange process. Collectively, these results strongly support the conclusion that PFOA displaces iodide in UiO-66-N(CH3)3+ due to its stronger and more competitive binding affinity for the quaternized ammonium groups. This anion exchange mechanism underpins the material's high selectivity and efficiency for PFOA adsorption.
Following ion exchange with SRB, the resulting composite, UiO-66-N(CH3)3+@SRB, exhibited a fluorescence “turn-off” state, marked by substantial quenching of SRB emission. This quenching effect prompted us to explore its potential as a “turn-on” fluorescence sensor for PFOA. Fluorescence emission spectra of the UiO-66-N(CH3)3+@SRB suspension in acetonitrile were recorded upon the incremental addition of minimal volume of aqueous stock solution of PFOA (0.5 mM). As shown in Fig. 5a, a clear fluorescence “turn-on” response was observed with increasing PFOA concentrations (0–5 μM). The obtained emission spectra closely resemble those of free SRB molecules in acetonitrile solution, confirming that SRB molecules are released into the solution upon adsorption of PFOA. In contrast, PFOS, one of the top two concerning PFAS compounds identified by the EPA, did not induce any fluorescence recovery (Fig. S19, ESI†). Notably, the system exhibited a 14-fold higher selectivity for PFOA over PFOS, despite their structural similarities.63 The fluorescence enhancement ratio (I/I0) at 574 nm reached up to 21, indicating a strong signal-to-background ratio. The fluorescence increase was also readily visible under UV illumination. The detection limit (LOD) for PFOA was determined to be 0.22 μM within the 0-3 μM dynamic concentration range, based on the standard IUPAC method (3σ) (Fig. 5b).
To further assess the sensor's selectivity, we evaluated its response to a range of potentially interfering analytes commonly present in aquatic environments, including lauric acid, octanoic acid, acetic acid, trifluoroacetic acid, sodium dodecyl sulfate (SDS), dodecane, phenol, and nitrobenzene. Interestingly, structurally similar non-fluorinated alkyl carboxylic acids such as lauric and octanoic acid did not induce a “turn-on” fluorescence response in the UiO-66-N(CH3)3+@SRB system (Fig. 5c and Fig. S20, ESI†). Similarly, smaller carboxylic acids (e.g., acetic and trifluoroacetic acid) and other small organic molecules (e.g., phenol and nitrobenzene) failed to generate any detectable fluorescence enhancement under the same conditions. Collectively, these results demonstrate that UiO-66-N(CH3)3+@SRB is highly selective for PFOA, supporting its potential as a sensitive and robust platform for detecting this high-priority PFAS compound identified by the EPA.
In addition to the fluorescence measurements presented in Fig. 5a, the IDA-based fluorescent sensor was further characterized using UV-vis absorption and NMR spectroscopy. Initially, UV-vis absorption spectra were measured for the centrifuged supernatant of UiO-66-N(CH3)3+@SRB suspensions before and after PFOA addition. Prior to adding PFOA, the supernatant exhibited no visible color or characteristic absorption peaks corresponding to free SRB molecules. However, following PFOA addition, the supernatant turned distinctly pink and showed a strong absorption peak at 553 nm, accompanied by a shoulder near 514 nm (Fig. 6). This absorption spectrum closely matched that of a standard SRB molecular solution, confirming that PFOA addition induces the release of SRB molecules into the solution.
![]() | ||
Fig. 6 UV-Vis absorption spectra of UiO-66-N(CH3)3+@SRB with and without PFOA. Inset shows the release of SRB dye into solution upon exposure to PFOA. |
The same supernatant samples obtained from the UiO-66-N(CH3)3+@SRB suspensions before and after PFOA addition were further analyzed using 1H-NMR spectroscopy. Prior to analysis, the supernatants were dried through solvent evaporation to isolate any residual compounds. The resulting residues were then dissolved in DMSO-d6 for 1H-NMR spectral analysis. As illustrated in Fig. S21 (ESI†), the presence of PFOA yielded prominent SRB-characteristic peaks at 8.2, 7.7, 7.1, and 6.9 ppm, whereas no significant signals appeared in the absence of PFOA. These findings, combined with the UV-vis and fluorescence spectral data, robustly confirm the IDA-based sensing mechanism. As outlined in Scheme 1, this mechanism involves the displacement-driven release of SRB molecules upon adsorption of PFOA, underscoring the sensor's effective and selective detection capability for PFOA.
This work further emphasizes the utility of post-synthetic modification in MOF design, enabling the development of advanced PFAS adsorbents and selective fluorescence probes. The findings presented here can be readily extended to other MOF systems, paving the way for future enhancements in adsorption capacity, kinetics, selectivity, and sensitivity. Efforts such as tailoring functional groups and crystal structures will contribute to the advancement of scalable, efficient technologies for water purification and environmental sensing.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5tc01765c |
This journal is © The Royal Society of Chemistry 2025 |