Full Paper www.afm-journal.de Molecular Tuning of Redox-Copolymers for Selective Electrochemical Remediation Kwiyong Kim, Paola Baldaguez Medina, Johannes Elbert, Emmanuel Kayiwa, Roland D. Cusick, Yujie Men, and Xiao Su* Molecular design of redox-materials provides a promising technique for tuning physicochemical properties which are critical for selective separations and environmental remediation. Here, the structural tuning of redox-copolymers, 4-methacryloyloxy-2,2,6,6-tetramethylpiperidin-1-oxyl (TMA) and 4-methacryloyloxy-2,2,6,6-tetramethylpiperidine (TMPMA), denoted as P(TMAx-co-TMPMA1−x), is investigated for the selective separation of anion contaminants ranging from perfluorinated substances to halogenated aromatic compounds. The amine functional groups provide high affinity toward anionic functionalities, while the redox-active nitroxyl radical groups promote electrochemically-controlled capture and release. Controlling the ratio of amines to nitroxyl radicals provides a pathway for tuning the redox-activity, hydrophobicity, and binding affinity of the copolymer, to synergistically enhance adsorption and regeneration. P(TMAx-co-TMPMA1−x) removes a model perfluorinated compound (perfluorooctanoic acid (PFOA)) with a high uptake capacity (>1000 mg g−1) and separation factors (500 vs chloride), and demonstrates exceptional removal efficiencies in diverse perand polyfluoroalkyl substances (PFAS) and halogenated aromatic compounds, in various water matrices. Integration with a boron-doped diamond electrode allows for tandem separation and destruction of pollutants within the same electrochemical cell, enabling the energy integration of the separation step with the catalytic degradation step. The study demonstrates for the first time the tuning of redox-copolymers for selective remediation of organic anions, and integration with an advanced electrochemical oxidation process for energy-efficient water purification. Dr. K. Kim, P. Baldaguez Medina, Dr. J. Elbert, Prof. X. Su Department of Chemical and Biomolecular Engineering University of Illinois at Urbana-Champaign Urbana, IL 61801, USA E-mail: x2su@illinois.edu E. Kayiwa, Prof. R. D. Cusick Department of Civil and Environmental Engineering University of Illinois at Urbana-Champaign Urbana, IL 61801, USA Prof. Y. Men Department of Chemical and Environmental Engineering University of California Riverside, CA 92521, USA The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/adfm.202004635. DOI: 10.1002/adfm.202004635 Adv. Funct. Mater. 2020, 30, 2004635 1. Introduction Achieving molecular selectivity is a critical challenge for separation processes, with advanced materials being key to enabling more efficient technologies in water purification and environmental remediation.[1] Selective separation of organic anions from industrial wastewater is of particular importance due to their environmental concern and dilute nature (micro to millimolar range); the presence of competing ions make it challenging to separate by conventional adsorption or ion-exchange methods.[2] There have been intense efforts for targeting a broad range of organic micropollutants,[3] but performance is often affected by intrinsic limitations in materials chemistry, as well as the high energy, operational, and chemical consumption costs.[4] Electrochemical separations provide modularity and scalability, with electrosorption methods allowing for the capture and release of target compounds based solely on potential control.[2,5] Recently, the focus of electrochemical separations has shifted from bulk capacitive desalination to ion-selective separation technologies.[1a,6] Redox-active materials provide an attractive electrosorption platform for the capture of specific pollutants, including heavy metal oxyanions, hydrophobic pollutants, and organic anions.[1a,2,7] Redox-electrodes have been studied also as catalytic interfaces for degradation of organic contaminants,[8] opening the door to the integration of separation and reaction steps.[5,7b] In addition, the asymmetric design of redox-electrodes can lead to dramatic increases in energy efficiency for water purification systems, by reducing operating voltages, and enabling complementary operations at the anode and the cathode.[5,7c,9] Rational tuning of material properties is a critical direction in the development of next-generation sorbents for anion micropollutants, especially the control of electrostatics, chargetransfer, and hydrophobic properties. Here, we demonstrate for the first time how molecularly designed copolymers can achieve highly selective separation of organic anions through electrochemical control, and ultimately integrate separation and remediation within a single electrochemical device, with 2004635 (1 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de enhanced energy efficiency. We explore polymer-functionalized electrodes bearing amine (NH) functionalities designed to provide high affinity for anionic organic compounds, and at the same time, control the loading of redox-groups, for example, nitroxide radicals (NO•), which can become charged to oxoammonium cation (+NO), for electrically-controlled capture and release. Amine functionalities have served as attractive binding sites for anionic functionalities by targeting the carboxylate and sulfonate head groups.[4b,10] We propose the use of 2,2,6,6-tetramethylpiperidine as a novel platform system, as it contains strong amine bases,[11] which can be protonated in a broad pH range, making these systems advantageous for adsorption of anionic functional groups. At the same time, controlled oxidation of NH to NO• provides a way of tuning hydrophobicity and redox-charge storage, thus assisting in the electrochemically-mediated capture and release. To molecularly optimize the various physico–chemical properties at play, we synthesized copolymers of 4-methacryloyloxy-2,2,6,6-tetramethylpiperidin-1-oxyl (TMA) and 4-methacryloyloxy-2,2,6,6-tetramethylpiperidine (TMPMA), namely, P(TMAx-co-TMPMA1−x), with x indicating the ratio of TMA in the copolymer (Figure 1a). Incorporating NO• radical moieties (increasing x) decreases the number of attractive NH binding sites (1−x), but at the same time increases hydrophobicity (NO• is more hydrophobic than NH) and reversible electrochemical activity (NO• is redox-active, while NH is not), which are all critical for synergistic binding and regeneration. Through mechanistic studies and separation tests, we demonstrate that control of the copolymer ratio leads to desirable tunability of redox-activity, hydrophobicity, and degree of NH affinity sites, thus enabling molecular selectivity for the separation of organic compounds ranging from persistent perfluorinated pollutants to halogenated aromatic contaminants. Per- and polyfluoroalkyl substances (PFAS) and halogenated compounds are particularly challenging environmental targets,[3] due to their strong chemical persistence and distinct hydrophobicity and charge properties. Unlike purely inorganic ions, these compounds possess surfactant-like properties (hydrophilic charged heads and hydrophobic tails), which offer unique challenges for selective interfacial binding. In our proposed copolymer system, the co-existing NH and NO• functionalities synergistically combine the affinity of NH sites with the electrostatic enhancement of the redox-active group, NO• (Figure 1b), promoting the electrochemically-controlled capture and release, without the need for chemical regenerants or harsh operational requirements (Figure 1c). Furthermore, we propose the first integration of polymer-functionalized electrode with a boron- Figure 1. a) Synthesis route of P(TMAx-co-TMPMA1−x). b) Redox reaction of redox-active TMA unit in P(TMAx-co-TMPMA1−x) (A denotes a generic anion). c) A schematic illustration showing electrochemically-controlled capture and release of PFOA by P(TMAx-co-TMPMA1−x). During adsorption, N-H sites of piperidine in TMPMA and redox-active oxoammonium in oxidized TMA units work in a synergistic way for enhanced electrostatic attraction of PFOA. During regeneration, redox-mediated charge repulsion enhanced by nitroxyl/oxoammonium couple facilitates electrochemically-controlled release. The ratio between TMA and TMPMA also tunes hydrophobicity, to achieve an optimal degree for PFOA binding and reversible release. Adv. Funct. Mater. 2020, 30, 2004635 2004635 (2 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de doped diamond (BDD) electrode to achieve separation and destruction of a model perfluorinated compound within an asymmetric electrochemical device; the coupled regeneration of P(TMAx-co-TMPMA1−x) with BDD allows for a modular, energy efficient, and selective reactive separation concept for organic anions. 2. Results and Discussion 2.1. Synthesis and Characterization of P(TMAx-co-TMPMA1−x) P(TMAx-co-TMPMA1−x) (with x indicating the ratio of TMA in the copolymer) was synthesized by partial oxidation of PTMPMA with m-chloroperbenzoic acid, as illustrated in Figure 1a. The detailed methodology can be found in Section S1.1, Supporting Information.[12] Synthetic control of the radical content (x) was confirmed through a number of analytical techniques. The presence of stable NO• radicals was investigated using ultraviolet-visible (UV–vis) analysis, where typical adsorption at 460 nm is associated with the n–π* transition of NO• radicals.[13] An increase in the concentration of NO• radicals, corresponding to an increase in degree of oxidation, was observed in the UV–vis spectra (Figure 2a). Quantification of radical yields based on TEMPO standards (Figure S4, Supporting Information) revealed 0%, 18%, 51%, 84% conversion of TMPMA to TMA at different degrees of oxidation. The same trend in radical densities were also monitored via electron paramagnetic resonance (EPR) (Figure S5, Supporting Information). EPR characterization of all P(TMAx-co-TMPMA1−x) samples showed broad singlets, in accordance with previous literature.[13] The signal intensity associated with the radical spin density increases with higher degree of oxidation. The oxidation of NH functionalities to NO• radicals was further confirmed via Fourier-transform infrared spectroscopy (FTIR) (Figure S6, Supporting Information). The FT-IR signals revealed that NO• in P(TMAx-co-TMPMA1−x) started to appear upon oxidation as evidenced by the appearance of NO• peak at 1467 cm−1.[12] At the same time, P(TMA84-co-TMPMA16), with the highest radical content, did not exhibit discernable peak of amine NH stretch in the range of 3400–3100 cm−1.[14] X-ray photoelectron spectroscopy (XPS) also indicated the conversion Figure 2. a) UV–vis spectra of P(TMAx-co-TMPMA1−x) copolymers in chloroform (concentration: 8 mg mL−1). b) Cyclic voltammograms of P(TMAx-coTMPMA1−x)-CNT electrodes in 0.1 m NaClO4 (scan rate: 10 mV s−1). c) PFOA uptake capacity and regeneration efficiency of P(TMAx-co-TMPMA1−x)-CNT electrodes (adsorption: in 20 mm NaCl + 0.1 mm (black) or 1 mm PFOA (dark gray and gray) at open circuit (O.C., black, dark gray) and +1.0 V (gray) for 0.5 h; and desorption (red) in 20 mm NaCl at −0.5 V for 0.5 h. Before desorption, adsorption was carried out at +1.0 V for 0.5 h in 1 mm PFOA + 20 mm NaCl (gray bars). In (a–c), x denotes the fraction of TMA in P(TMA-co-TMPMA). d) PFOA uptake capacity of the P(TMA51-co-TMPMA49)-CNT electrodes at various polarization conditions in 0.1 mm PFOA + 20 mm NaCl for 0.5 h. e) Images of water droplet on the surface of P(TMAx-co-TMPMA1−x)-coated electrodes in contact angle measurement (contact angles were 70°, 94°, and 126° for 0%, 51%, and 51% after adsorption, respectively). f) Comparison of PFOA uptake capacity of CNT and P(TMA51-co-TMPMA49)-CNT electrodes in 0.1 mm PFOA + 20 mm NaCl for 0.5 h. g) Scanning electron microscopy (SEM) of P(TMA51-co-TMPMA49)-CNT electrodes, and EDS mapping of nitrogen and adsorbed fluoride. Scale bar for high-resolution SEM is 2 µm, and scale bar for EDS and N/F mapping is 200 µm. Adv. Funct. Mater. 2020, 30, 2004635 2004635 (3 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de of NH functionalities to NO• radicals (Figure S7, Supporting Information). The peak at the lower binding energy of 398.9–399.7 eV is assigned to NH groups, and the other peak at about 401.3–401.5 eV is the characteristic peak of NO• radicals.[12,15] When the relative concentration of each component representing NH and NO• is compared, the same trend of increasing NO• content could be observed with higher degree of oxidation. The NO• contents obtained from XPS exhibited lower values compared to UV–vis-based quantification of NO•, and this can be ascribed to the distribution of more polar NH group on the surface caused during the precipitation of polymers in methanol or water (see Section S1.1, Supporting Information). Finally, a higher electrochemical activity with higher x in P(TMAx-co-TMPMA1−x) was clearly observed using cyclic voltammetry (CV), measured using heterogeneous P(TMAx-coTMPMA1−x) interfaces blended with CNT as a conductive additive formed by dip-coating method (Figure 2b).[5] The P(TMAx-co-TMPMA1−x)-CNT electrode showed homogeneous and nanoporous morphology (Figure 2g), with the CNTs providing a porous network for ion accessibility. The reversible peaks in Figure 2b could be ascribed to the one-electron redox reaction between NO• and +NO, shown in Figure 1b.[13] The normalized current increased proportionally to the increase in oxidation degree, indicating the larger number of reversible redox couples of NO• and +NO in P(TMAx-co-TMPMA1−x) at higher x. However, the peak separation became larger with higher oxidation degree, probably due to increased hydrophobicity from the replacement of more hydrophilic NH with less hydrophilic NO•, thus limiting ion accessibility. The spectroscopic and electrochemical characterizations confirm the controllable composition of NO• and NH loadings through the synthesis of P(TMAx-co-TMPMA1−x) copolymer, and tunable electrochemical activity. 2.2. Selective (Electro)Sorption of a Model Compound PFOA The separation performance of P(TMAx-co-TMPMA1−x)-CNT electrodes, for each radical content x, were first evaluated by carrying out the electrosorption of the model compound PFOA at initial concentrations of 0.1 and 1 mm in the presence of 20 mm NaCl (Figure 2c). These initial concentrations translate to a molar ratio of PFOA to total nitrogen species (sum of NH and NO•) in P(TMAx-co-TMPMA1−x) of 0.5:1 and 5:1, respectively. First, the highest adsorption uptake of PFOA in the absence of electricity input (at open circuit) was obtained with polymers having a lower content of nitroxide radicals: the non-oxidized polymer (x = 0%) showed the highest uptake at 0.1 mm (299.6 mg g−1) and the polymer with x = 18% had the highest uptake at 1 mm (1199.3 mg g−1) (Figure 2c). The trend of decreasing PFOA removal by polymers with >18% radical content indicates that the piperidine groups are essential for high PFOA binding when electricity is not applied. At the initial PFOA concentration of 1 mm (dark gray bars in Figure 2c), the molar ratio of adsorbed PFOA to amine functional groups in P(TMAx-co-TMPMA1−x) (PFOA/NH) was 0.48:1 at x = 0%. The incomplete utilization of NH, even in the presence of high initial PFOA concentration (1 mm), could be ascribed to Adv. Funct. Mater. 2020, 30, 2004635 the hydrophilic nature of the piperidine groups (Figure 2e). Incorporating NO• radical moieties increase the hydrophobicity, and as a result, the utilization of NH functionalities approached a stoichiometric ratio of 1:1 to PFOA at x = 18% and 51%. The close to full utilization of the NH functionalities indicates that piperidine groups are critical to driving the selective binding of PFOA under open circuit condition. Interestingly, applying an increasingly positive potential beyond open-circuit improved the adsorption of PFOA onto the copolymer electrodes. For example, when the applied potential at P(TMA51-co-TMPMA49) was finely controlled in 0.1 mm PFOA + 20 mm NaCl, a positive correlation between applied potential and uptake of PFOA was observed (Figure 2d). In the same electrolyte condition (0.1 mm PFOA + 20 mm NaCl), the potential at open circuit was stabilized at −0.28 ± 0.04 V, and applied potentials higher than that exhibited better uptake compared to open circuit adsorption (Figure 2d). In Figure 2c, when x was varied, applying a positive bias of +1.0 V (vs Ag/ AgCl) enhances PFOA uptake compared to open circuit, showing >1000 mg PFOA/g adsorbent for all cases at 1 mm PFOA, with higher uptake being observed at lower x. In particular, the molar ratio of adsorbed PFOA to NH (PFOA/ NH) in P(TMA84-co-TMPMA16) was 3.85:1 during the electrosorption at +1.0 V, indicating that PFOA binds not only onto NH, but also onto redox-active +NO sites via electrostatic attraction (see the anion doping in Figure 1b). After electrosorption, the increase in PFOA content on the copolymer film was evident through EDS mapping (Figure 2g), and both survey and high-resolution XPS analysis (Figures S9,S10, Supporting Information) indicated an increase in fluorine content. The contact angle analysis also confirmed the increase in hydrophobicity at the adsorbent layer after the electrosorption of PFOA (Figure 2e). Our result indicates the synergistic interplay between NH affinity sites and redox-promoted electrostatic interaction at the charged +NO sites for selective PFOA binding. The unfunctionalized CNT does not present any noticeable uptake, either under applied potential (36.3 mg g−1 at +1.0 V) or at O.C. (10.9 mg g−1) (Figure 2f), indicating that selective copolymer functionalization is responsible for significant uptake (30-fold increase with copolymer). Thus, all these results provide ample evidence that surface functionalization with P(TMAx-co-TMPMA1−x)-CNT enables molecular selectivity for PFOA uptake in the presence of background anions. Here, we define the separation factor of PFOA versus chloride as follows Separation factorPFOA − /Cl − = qPFOA − /qCl − (cPFOA − )0 /(c Cl − )0 (1) where (cPFOA−)0 and (cCl−)0 are the initial concentration in the solution, and qPFOA- and qCl− are the solid-phase uptakes of PFOA− and Cl−, the latter ratio which can be estimated from XPS analysis (see Section S1.4, Supporting Information, for calculation details). For example, at x = 51%, in the presence of 200-fold excess of chloride (0.1 mm PFOA + 20 mm NaCl), a molar ratio comparison revealed the PFOA− to Cl− ratio to be 2.5 at open circuit, equivalent to a separation factor of 500 (Figure 3a). If final equilibrium concentrations were used to estimate the separation factor,[1a] then the separation factor 2004635 (4 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de Figure 3. a) Separation factors of PFOA− versus chloride on P(TMA51-co-TMPMA49)-CNT electrodes after adsorption in 0.1 mm PFOA + 20 mm NaCl for open circuit, and different positive potentials versus Ag/AgCl. b) Adsorption kinetics of PFOA by P(TMA51-co-TMPMA49)-CNT electrode at open circuit and +1.0 V (in 0.1 mm PFOA + 20 mm NaCl). c) Equilibrium isotherm of P(TMA51-co-TMPMA49)-CNT electrode at open circuit for a range of PFOA concentrations in 20 mm NaCl (blue line represents Freundlich isotherm). d) PFOA uptake capacity (blue dots) and regeneration efficiency (red bars) of P(TMA51-co-TMPMA49)-CNT electrode over 5 cycles (adsorption: 0.1 mm PFOA + 20 mm NaCl at +1.0 V for 0.5 h; desorption: in 20 mm NaCl at −1.0 V for 1 h). toward PFOA over chloride would be 832 at these conditions. With the positive potential at +1.0 V, the PFOA− to Cl− ratio was decreased to 0.76, but the separation factor was 152 due to the much lower concentration of PFOA compared to chloride, indicating an exceptionally strong interaction between PFOA and the copolymer (Figure 3a). Applying more positive potential resulted in more competitive attraction of chloride, thereby leading to decreasing trend in Coulombic efficiency for PFOA adsorption (Figure S11, Supporting Information) and also in separation factor (Figure 3a, 37.4 at +2.5 V). This indicates that a certain potential window (<1.2 V vs Ag/AgCl) is recommended to achieve higher molecular selectivity (separation factor > 100). A lower potential window also minimizes energy consumption—as will be discussed later, a practical operational potential range could be from +0.6 to +1.2 V versus Ag/AgCl at the working copolymer electrode. 2.3. Redox-Mediated Regeneration Here, we investigated the effect of redox-active NO• functional groups on promoting electromediated release of the bound PFOA species. Redox-active species have garnered intense attention for energy storage,[2,16] and more recently, have been show to efficiently couple charge storage with electrochemically-switched capture and release of charged pollutants.[5,7b] As shown in XPS analysis (Figure S8, Supporting Information) and as observed in literature,[17] one-electron oxidation of NO• radicals forms +NO, and its fully reversible redox process (Figure 2b) accompanies the uptake or expulsion of doped anionic species for charge neutrality (Figure 1b). We investigated the effect of the oxidation degree and copolymer structure on the regeneration efficiency. For each different copolymer Adv. Funct. Mater. 2020, 30, 2004635 (varying x), we first adsorbed the PFOA at +1.0 V versus Ag/ AgCl, and then applied −0.5 V versus Ag/AgCl to release the adsorbed molecules. As depicted in Figure 2c, the regeneration efficiency (defined as the ratio of PFOA recovered to adsorbed) improved as the radical yield was increased from x = 0% to 51%. This result indicated that anion release induced by redox-mediated charge repulsion (as shown in Figure 1b) is more efficient than discharging based solely on polarization of a conductive surface. For x = 51%, the regeneration efficiency was enhanced by applying a more negative bias, exhibiting >80% regeneration efficiency at −1.5 V (Figure S12, Supporting Information). It is interesting to note that P(TMA84-co-TMPMA16), despite having the highest oxidation degree, showed the poorest regeneration. This could be ascribed to the background hydrophobic interaction between PFOA and neutral, hydrophobic PTMA layer, which is supported by contact angle analysis indicating more hydrophobic features at higher NO• content (Figure 2e). Our results confirmed that compositional control of the copolymer, with delicately tuned NH content, redox-activity, and hydrophobicity, enables not only significant uptake of PFOA but also efficient electrochemically-mediated release of adsorbed PFAS molecules. Through structural and electrochemical control of the redox-copolymer electrodes, we enable reuse of the adsorbent without the need for chemical regenerants and/or harsh operational environments. 2.4. Adsorption Kinetics and Stability Based on the molecular optimization showing that the highest regeneration efficiency occurs at x = 51%, P(TMA51co-TMPMA49) was selected for the subsequent experiments on adsorption kinetics, isotherm, and cyclability. Adsorption 2004635 (5 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de kinetics were studied both at open circuit and +1.0 V in 0.1 mm PFOA + 20 mm NaCl. As seen in Figure 3b, 81% of removal was observed within 1 h by applying +1.0 V. The kinetics in the first 3 h at +1.0 V outperformed open circuit, in accordance with our previous observations about the contribution of the radical-based electrostatic interaction to enhancing PFOA binding (Figure 2c,d). However, the maximum obtainable uptake after reaching equilibrium (>10 h) was slightly higher at open circuit compared to positive bias. It is hypothesized that electrostatic interactions play a significant role of accelerating initial kinetics, but prolonged polarization induced competitive adsorption of chloride, as reflected in the separation factor and Coulombic efficiency analysis (Figure 3a; Figure S11, Supporting Information). Thus, in practice, electrosorption duration can be controlled to optimize separation factor. Equilibrium isotherm studies show that a maximum capacity of 970 mg PFOA/g adsorbent can be achieved at open circuit (Figure 3c), which is comparable or higher than reported loadings for various chemical adsorbents for PFOA (Table S4, Supporting Information), except here we are highly selective and electrochemically-mediated. A cycling test was conducted with the P(TMA51-co-TMPMA49)CNT electrode being charged at +1.0 V, in the presence of 0.1 mm PFOA and 20 mm NaCl for 0.5 h, and then discharged at −1.0 V into an NaCl solution for 1 h. Figure 3d shows the regeneration efficiency was maintained at >90% for a number of cycles per step starting at cycle 2. The incomplete regeneration at the initial release cycle 1 can be ascribed to initial irreversible adsorption due to hydrophobic interactions between the discharged (neutral) polymer and PFOA. Even so, the working capacity of the functionalized polymer electrode was preserved at close to 200 mg PFOA/g adsorbent for a number of cycles. This proves the reusability advantages of our P(TMAx-co-TMPMA1−x)-CNT electrodes, which are fully electrochemically-controlled and thus do not require chemical regenerants for PFOA release. 2.5. Real Water Matrices and Separation of Contaminants with Diverse Chemical Structures To evaluate relevant environmental conditions, the removal of PFOA was tested under different water matrices (Figure 4a), including 20 mm NaCl, tap water, 5 mm Na2CO3, and municipal secondary wastewater effluent solution (obtained from Urbana-Champaign Sanitary District). In all water matrices, >93% removal efficiency was achieved within 3 h of electrosorption when spiked with 0.1 mm PFOA, and >82.5% at the dilute initial PFOA concentration of 100 ppb (0.242 µm) to simulate environmental conditions. We also tested the performance of P(TMA51-co-TMPMA49)-functionalized electrodes for the removal of several PFAS compounds with different chemical structures, including perfluorocarboxylic acids (CnF2n+1COO−), per- and polyfluoro dicarboxylic acids (−OOC−CnF2n−COO−), and perfluoroalkanesulfonic acids (CnF2n+1SO3−) (Figure 4b). For a given chain length (n = 8), the open circuit adsorption of different PFAS compounds seemed to be mainly affected by hydrophobic interaction; for example, PFSA (perfluorosebacic acid), having double COO− head groups and lacking primary CF bond, is less hydrophobic and thus shows the poorest uptake at open circuit. For all cases, however, applying +1.0 V of positive potential led to greatly improved uptake, with PFSA having the highest increase with the positive potential. The effect of the chemical structure of PFAS on the electrosorption provides further supporting evidence that electrostatic interaction plays a dominant mechanism to balance the affinity interaction, and that molecular control of the copolymer ratio is thus critical to maximizing these selective interactions. The regeneration efficiencies of P(TMA51-co-TMPMA49)-CNT with different PFAS compounds seemed to be also affected by the hydrophobicity of PFAS compounds (Figure S15, Supporting Information); having longer chain length showed decreased regeneration. At the same time, the regeneration was facilitated in the case of PFSA, which has two COO− head groups. Thus, molecular tuning of copolymer ratio is shown to optimize regeneration for various PFAS compounds. Furthermore, the versatility of our redox-copolymer system was demonstrated by the effective separation of halogenated aromatic compounds, including 2,4-dichlorobenzoic acid, ibuprofen, and clofibric acid (Figure S14, Supporting Information). All these results support the wide applicability of our molecularly-tuned electroactive copolymers electrode to address real-world contamination under relevant conditions, and a diversity of chemical structures. Figure 4. a) Removal of PFOA in different water matrices spiked with 0.1 mm or 100 ppb (0.242 µm). +1.0 V was applied for 3 h. b) Uptake capacity of different PFAS compounds with P(TMA51-co-TMPMA49)-CNT electrode at open circuit and +1.0 V (adsorption: 0.1 mm PFAS compounds + 20 mm NaCl for 0.5 h). Adv. Funct. Mater. 2020, 30, 2004635 2004635 (6 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de Figure 5. a) Schematic illustrations showing selective uptake of PFOA by P(TMAx-co-TMPMA1−x)-CNT during adsorption and degradation of released PFOA by BDD electrode during regeneration. b) Simultaneous tracking of anode (BDD) and counter electrode potentials during chronopotentiometry at 10 mA cm−2 in 0.1 mm PFOA + 20 mm NaCl for 2 h using different counter electrodes of Pt, CNT, and P(TMA51-co-TMPMA49)-CNT. (c) and (d) represent the operation of the asymmetric P(TMA51-co-TMPMA49)-CNT//BDD system for one cycle of adsorption and desorption (as depicted in [a]): c) PFOA uptake capacity during the adsorption and defluorination efficiency during the regeneration in the asymmetric P(TMA51-co-TMPMA49)-CNT// BDD configuration (adsorption: in 0.1 mm PFOA + 20 mm NaCl for 0.5 h). d) Tracking of anode (BDD) and cathode (P(TMA51-co-TMPMA49)-CNT) potentials during the regeneration (desorption) of P(TMA51-co-TMPMA49)-CNT in 20 mm NaCl at 10 mA cm−2 for 5 h. 2.6. Integration of Reactive Separation While the molecular control of the redox-copolymer allows for the selective separation of organic contaminants, the fate of the enriched pollutant waste after separation process must be addressed.[18] Here, we propose a unique approach to achieve tandem separation and degradation of the model compound PFOA within the same electrochemical unit, by leveraging an asymmetric electrochemical design that combines the P(TMAxco-TMPMA1−x)-functionalized CNT working electrode with a BDD counter electrode. While the redox-copolymer acts as a selective adsorbent, the BDD counter-electrode acts as a synergistic electrocatalyst during the regeneration step. As shown in Figure 5a, PFOA can be first electrochemically captured by P(TMAx-co-TMPMA1−x), whereas during release, BDD electrode catalyzes defluorination of released PFOA. BDD anode has proven its ability to degrade PFAS and various organic contaminants in wide applications for electrochemical advanced oxidation, showing desirable features of commercial availability, high stability, low adsorption capacity, and high overpotential for oxygen evolution (Figure S3, Supporting Information).[19] First, chronopotentiometric tests on the BDD anode were carried out at 10 mA cm−2 with various counter electrodes of platinum, CNT, and P(TMA51-co-TMPMA49)-CNT in 0.1 mm PFOA + 20 mm NaCl; simultaneous tracking of the potential at both working (BDD) and counter electrodes during charging were carried out against a reference Ag/AgCl (Figure 5b). The analysis showed that conductive, non-redox counter electrodes (platinum and Adv. Funct. Mater. 2020, 30, 2004635 CNT) reached significantly negative overpotentials; while the redox-active P(TMA51-co-TMPMA49)-CNT on the other hand remains at potentials no lower than −1.4 V (Figure 5b). Thus, our P(TMA51-co-TMPMA49) provides an energy-efficient option as a counter electrode during electrochemical degradation, enabling much lower voltage windows for BDD-based electrochemical oxidation processes. On the conductive counter electrodes, water decomposition can be a major cathodic parasitic reaction at very negative overpotentials.[5,7c] By coupling the P(TMA51-coTMPMA49)-CNT with BDD, we can efficiently store electrons through NO•/+NO electrochemistry at more moderate overpotentials to avoid water splitting, and while its redox-process can be directly connected to electrochemically-controlled PFOA capture and release. Next, tandem separation and degradation of PFOA within a single device (Figure 5a) was experimentally realized using an asymmetric configuration of P(TMA51-co-TMPMA49)-CNT// BDD. First, we carried out electrosorption in a 0.1 mm PFOA + 20 mm NaCl for 0.5 h at +1.0 V, followed by applying constant current density of 10 mA cm−2 with reversed polarity for 5 h into a 20 mm NaCl solution to release and degrade PFOA. The adsorption capacity was 291 mg PFOA/g adsorbent (Figure 5c), and applying constant current density of 10 mA cm−2 achieved a high regeneration of the bound PFOA, while pushing the potential at the P(TMA51-co-TMPMA49)-CNT electrode to be −1.2 V (Figure 5d). At the same time, the working potential of BDD anode was maintained at about +4.5 V (Figure 5d), a potential range capable of promoting the degradation of PFOA 2004635 (7 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de Figure 6. a) Energy consumption during the electroseparation of PFOA by CNT and by P(TMAx-co-TMPMA1−x)-CNT. b) PFOA removal (sky blue bar), defluorination efficiency (orange bar), and energy consumption (brown bar) during chronopotentiometry at 10 mA cm−2 in 0.1 mm PFOA + 20 mm NaCl for 2 h using BDD as a working electrode and different counter electrodes of Pt, CNT, and P(TMA51-co-TMPMA49)-CNT. via electron transfer from the head group to the BDD.[19a] From LC/MS analysis, no PFOA was found in the desorption electrolyte, implying complete degradation of released PFOA. The final concentration of fluoride was quantified using ion chromatography, and the normalized fluoride recovery out of adsorbed amount (291 mg PFOA/g adsorbent) was found to be 81% after 5 h electrolysis (Figure 5c), which is comparable or higher than the defluorination ratio (i.e., CF− produced/CF− in PFOA,initial) reported for various electrochemical PFOA degradation.[19d,20] Its corresponding F index (i.e., CF− produced/CPFOA degraded) was 12.1, implying fluoride recovery of 12.1 out of 15 per PFOA molecule degraded. To the best of our knowledge, our proofof-concept study is the first realization of the coupling of a redox-active polymer with BDD electrode, providing an efficient remediation pathway for separating and degrading enriched PFAS compounds within a single unit operation. 2.7. Energy Analysis Finally, the energy analysis for electrochemical separation and defluorination highlighted the intrinsic performance advantages of our molecularly-engineered copolymer over regular capacitive electrosorption systems. First, we achieved remarkable improvements in energy-efficiency in the electroseparation process. Electrosorption using P(TMAx-co-TMPMA1−x)-CNT enabled a significantly improved separation factor, leading to high adsorption capacity and a lowering of the energy input in the separation process down to <0.44 kWh mol−1 PFOA (Figure 6a). On the other hand, control experiments of PFOA electrosorption on CNT electrodes was estimated to consume >2.67 kWh mol−1 PFOA (being potentially as high as 196 kWh mol−1 PFOA), due to low selectivity and poor affinity of the non-functionalized surfaces (Figure 6a). Furthermore, the polymer-functionalized interfaces, when coupled with BDD anode, provided a lower energy consumption via tuning of the redox potentials. In chronopotentiometric operations for 2 h with three different counter electrodes of platinum, CNT, and P(TMA51-co-TMPMA49)-CNT, all the configurations exhibited similar PFOA removal and defluorination performance (Figure 6b). However, a much lower mineralization energy was required with redox-active copolymer counter-electrodes Adv. Funct. Mater. 2020, 30, 2004635 (P(TMA51-co-TMPMA49)-CNT) compared to conductive ones (CNT and platinum) (Figure 6b). For an adsorption and regeneration cycle of the P(TMA51-coTMPMA49)-CNT//BDD configuration (Figure 5a), the mineralization energy for PFOA estimated from the defluorination performance (Figure 5c) and voltage/current profile (Figure 5d) was found to be 62.3 kWh mol−1 F− for a 5 h operation. This energy consumption is among the lowest reported in the literature (prior electrochemical [10–1500 kWh mol−1 F−], microwave [4200 kWh mol−1 F−], UV [450–2000 kWh mol−1 F−], and ultrasonic [250–4200 kWh mol−1 F−]).[20b,21] The significantly lower energy consumption can be attributed to the high defluorination efficiency, combined with the tuning of redox-potential through the P(TMA51-co-TMPMA49)-CNT electrode. Our consideration in the energy aspect demonstrates the possibility of using P(TMAx-co-TMPMA1−x)-CNT as not only a cost-effective electrochemically-mediated adsorbent, but also a sustainable counter electrode enabling energy-efficient, tandem degradation of PFOA, providing a unique pathway for next-generation, energy-integrated water purification devices. 3. Conclusions We have developed a copolymer of 4-methacryloyloxy-2,2,6,6tetramethylpiperidin-1-oxyl (TMA) and 4-methacryloyloxy2,2,6,6-tetramethylpiperidine (TMPMA), namely, P(TMAx-coTMPMA1−x), for the electrochemically-controlled capture and release of organic contaminants. Control of the radical yields (x) within the copolymer tunes the affinity toward organic anions, and the optimal ratio between NH and NO• functionalities allows for a selective and reversible electrochemical adsorption. P(TMAx-co-TMPMA1−x)-CNT is shown to be effective for selective capture of the model compound PFOA in the presence of excess competing ions, with desirable features of high adsorption capacity and cyclability. The performance of P(TMAx-co-TMPMA1−x)-CNT was benchmarked using various real-world water matrices spiked 0.1 mm and 100 ppb of PFOA, and also using various PFAS with differing chain lengths and chemical end-groups, and further expanded for the separation of halogenated aromatic contaminants. The P(TMAx-coTMPMA1−x)-CNT adsorbent electrode was integrated with a 2004635 (8 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de BDD counter-electrode to establish a unique asymmetric configuration. The combined system coupled selective separation by P(TMAx-co-TMPMA1−x)-CNT with the simultaneous degradation by BDD during release of adsorbed PFOA, exhibiting a high defluorination performance with improved energy efficiency. Our work emphasizes the importance of judicious selection and tuning of the copolymer chemistry for electrochemicallymediated separation of organic micropollutants—by controlling the copolymer composition, we can dial-in desired properties such as binding affinity, electrostatics, and redox-activity. Despite the ultra-dilute concentration of organic anions and coexistence of competing constituents, molecular design of ionselective redox-active interfaces enables an exceptional binding affinity, with a high separation factor, and significantly enhanced regeneration by electrochemical control. We also demonstrated an energy-efficient and modular approach for asymmetrically integrating redox-active interfaces with electrocatalytic BDD, for the reactive separation of highly-stable fluorinated compounds. The proposed approach can significantly impact future electrochemical applications for water purification. Finally, we expect continued advances in the design of redox-materials to enable molecularly-integrated systems which not only perform selective separations, but also provide direct coupling with electrochemical energy storage and generation devices. Supporting Information Supporting Information is available from the Wiley Online Library or from the author. Acknowledgements K.K. and P.B.M. contributed equally to this work. The authors acknowledge the support of the University of Illinois, UrbanaChampaign (UIUC) for startup funding, the support of the National Science Foundation under Grant #1931941, the support of SURGE Fellowship from the UIUC, and partial funding under the Illinois Water Resources Center (IWRC) Annual grant (2019). SEM, XPS, FT-IR, GPC, and contact angle analysis were carried in the Frederick Seitz Materials Research Laboratory Central Research Facilities, University of Illinois. LC-MS and EPR were carried out in the School of Chemical Science (SCS) Mass Spectrometry Lab and EPR Lab, respectively. Major funding for the Bruker EMXPlus was provided by National Science Foundation Award 1726244 (2017) to the School of Chemical Sciences EPR lab at the University of Illinois. The authors thank Furong Sun and Dr. Toby Woods for their assistance with analysis. Conflict of Interest The authors declare no conflict of interest. Keywords electrocatalysis, electrochemical separation, molecular selectivity, redoxactive polymers, tunability Received: May 30, 2020 Revised: August 22, 2020 Published online: September 16, 2020 Adv. Funct. Mater. 2020, 30, 2004635 [1] a) P. Srimuk, X. Su, J. Yoon, D. Aurbach, V. Presser, Nat. Rev. Mater. 2020, 5, 517; b) H. Wang, X. Mi, Y. Li, S. Zhan, Adv. Mater. 2020, 32, 1806843; c) D. Pakulski, W. Czepa, S. D. Buffa, A. Ciesielski, P. Samorì, Adv. Funct. Mater. 2019, 30, 2070010; d) H. Qiu, M. Xue, C. Shen, Z. Zhang, W. Guo, Adv. Mater. 2019, 31, 1803772; e) Y. Han, Z. Xu, C. Gao, Adv. Funct. Mater. 2013, 23, 3693; f) J. S. Kang, S. Kim, D. Y. Chung, Y. J. Son, K. Jo, X. Su, M. J. Lee, H. Joo, T. A. Hatton, J. Yoon, Y.-E. Sung, Adv. Funct. Mater. 2019, 30, 1909387. [2] X. Su, H. J. Kulik, T. F. Jamison, T. A. Hatton, Adv. Funct. Mater. 2016, 26, 3394. [3] a) A. Alsbaiee, B. J. Smith, L. Xiao, Y. Ling, D. E. Helbling, W. R. Dichtel, Nature 2016, 529, 190; b) E. Gagliano, M. Sgroi, P. P. Falciglia, F. G. A. Vagliasindi, P. Roccaro, Water Res. 2020, 171, 115381; c) T. D. Appleman, C. P. Higgins, O. Quiñones, B. J. Vanderford, C. Kolstad, J. C. Zeigler-Holady, E. R. V. Dickenson, Water Res. 2014, 51, 246. [4] a) Z. Wei, T. Xu, D. Zhao, Environ. Sci.: Water Res. Technol. 2019, 5, 1814; b) M. Ateia, A. Alsbaiee, T. Karanfil, W. Dichtel, Environ. Sci. Technol. Lett. 2019, 6, 688. [5] K. Kim, S. Cotty, J. Elbert, R. Chen, C. H. Hou, X. Su, Adv. Mater. 2020, 32, 1906877. [6] a) S. K. Patel, C. L. Ritt, A. Deshmukh, Z. Wang, M. Qin, R. Epsztein, M. Elimelech, Energy Environ. Sci. 2020, 13, 1694; b) X. Su, Curr. Opin. Colloid Interface Sci. 2020, 46, 77; c) X. Su, T. A. Hatton, Phys. Chem. Chem. Phys. 2017, 19, 23570. [7] a) R. Chen, T. Sheehan, J. L. Ng, M. Brucks, X. Su, Environ. Sci.: Water Res. Technol. 2020, 6, 258; b) X. Su, A. Kushima, C. Halliday, J. Zhou, J. Li, T. A. Hatton, Nat. Commun. 2018, 9, 4701; c) X. Su, K.-J. Tan, J. Elbert, C. Rüttiger, M. Gallei, T. F. Jamison, T. A. Hatton, Energy Environ. Sci. 2017, 10, 1272; d) X. Mao, W. Tian, Y. Ren, D. Chen, S. E. Curtis, M. T. Buss, G. C. Rutledge, T. A. Hatton, Energy Environ. Sci. 2018, 11, 2954. [8] X. Su, L. Bromberg, K.-J. Tan, T. F. Jamison, L. P. Padhye, T. A. Hatton, Environ. Sci. Technol. Lett. 2017, 4, 161. [9] K. J. Tan, X. Su, T. A. Hatton, Adv. Funct. Mater. 2020, 30, 1910363. [10] a) W. Ji, L. Xiao, Y. Ling, C. Ching, M. Matsumoto, R. P. Bisbey, D. E. Helbling, W. R. Dichtel, J. Am. Chem. Soc. 2018, 140, 12677; b) M. J. Klemes, Y. Ling, C. Ching, C. Wu, L. Xiao, D. E. Helbling, W. R. Dichtel, Angew. Chem., Int. Ed. 2019, 58, 12049; c) J. Li, Q. Li, L.-s. Li, L. Xu, Chem. Eng. J. 2017, 320, 501; d) D. Q. Zhang, W. L. Zhang, Y. N. Liang, Sci. Total Environ. 2019, 694, 133606. [11] a) I. Y. Zhukova, E. N. Papina, I. N. Tyaglivaya, IOP Conf. Ser.: Earth Environ. Sci. 2020, 459, 032011; b) J.-L. Roubaty, M. Bréant, M. Lavergne, A. Revillon, Die Makromol. Chem. 1978, 179, 1151. [12] L. Rostro, S. H. Wong, B. W. Boudouris, Macromolecules 2014, 47, 3713. [13] K. Zhang, Y. Hu, L. Wang, J. Fan, M. J. Monteiro, Z. Jia, Polym. Chem. 2017, 8, 1815. [14] M. A. Chinelatto, J. A. M. Agnelli, S. V. Canevarolo, Polím.: Ciênc. Tecnol. 2014, 24, 30. [15] M.-K. Hung, Y.-H. Wang, C.-H. Lin, H.-C. Lin, J.-T. Lee, J. Mater. Chem. 2012, 22, 1570. [16] a) J. Lee, P. Srimuk, S. Fleischmann, X. Su, T. A. Hatton, V. Presser, Prog. Mater. Sci. 2019, 101, 46; b) X. Su, T. A. Hatton, Adv. Colloid Interface Sci. 2017, 244, 6. [17] S. Wang, F. Li, A. D. Easley, J. L. Lutkenhaus, Nat. Mater. 2019, 18, 69. [18] a) M. J. Bentel, Y. Yu, L. Xu, H. Kwon, Z. Li, B. M. Wong, Y. Men, J. Liu, Environ. Sci. Technol. 2020, 54, 2489; b) M. J. Bentel, Y. Yu, L. Xu, Z. Li, B. M. Wong, Y. Men, J. Liu, Environ. Sci. Technol. 2019, 53, 3718. [19] a) Q. Zhuo, S. Deng, B. Yang, J. Huang, B. Wang, T. Zhang, G. Yu, Electrochim. Acta 2012, 77, 17; b) B. Gomez-Ruiz, S. Gómez-Lavín, N. Diban, V. Boiteux, A. Colin, X. Dauchy, A. Urtiaga, Chem. Eng. J. 2017, 322, 196; c) Y. Juang, E. Nurhayati, C. Huang, J. R. Pan, S. Huang, Sep. Purif. Technol. 2013, 120, 289; d) C. E. Schaefer, 2004635 (9 of 10) © 2020 Wiley-VCH GmbH www.advancedsciencenews.com www.afm-journal.de C. Andaya, A. Burant, C. W. Condee, A. Urtiaga, T. J. Strathmann, C. P. Higgins, Chem. Eng. J. 2017, 317, 424. [20] a) Q. Zhuo, S. Deng, B. Yang, J. Huang, G. Yu, Environ. Sci. Technol. 2011, 45, 2973; b) T. X. H. Le, H. Haflich, A. D. Shah, B. P. Chaplin, Environ. Sci. Technol. Lett. 2019, 6, 504; c) C. E. Schaefer, C. Andaya, A. Urtiaga, E. R. McKenzie, C. P. Higgins, J. Hazard. Mater. 2015, 295, 170; d) B. Yang, C. Jiang, G. Yu, Q. Zhuo, S. Deng, J. Wu, Adv. Funct. Mater. 2020, 30, 2004635 H. Zhang, J. Hazard. Mater. 2015, 299, 417; e) Q. Ma, L. Liu, W. Cui, R. Li, T. Song, Z. Cui, RSC Adv. 2015, 5, 84856; f) H. Lin, J. Niu, S. Ding, L. Zhang, Water Res. 2012, 46, 2281; g) J. Niu, H. Lin, J. Xu, H. Wu, Y. Li, Environ. Sci. Technol. 2012, 46, 10191. [21] a) Y. B. Hu, S. L. Lo, Y. F. Li, Y. C. Lee, M. J. Chen, J. C. Lin, Water Res. 2018, 140, 148; b) Z. Xu, Y. Yu, H. Liu, J. Niu, Sci. Total Environ. 2017, 579, 1600. 2004635 (10 of 10) © 2020 Wiley-VCH GmbH