ACS Publications. Most Trusted. Most Cited. Most Read
Cross-Linking of Sugar-Derived Polyethers and Boronic Acids for Renewable, Self-Healing, and Single-Ion Conducting Organogel Polymer Electrolytes
My Activity

Figure 1Loading Img
  • Open Access
Article

Cross-Linking of Sugar-Derived Polyethers and Boronic Acids for Renewable, Self-Healing, and Single-Ion Conducting Organogel Polymer Electrolytes
Click to copy article linkArticle link copied!

  • Emma L. Daniels
    Emma L. Daniels
    University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.
    Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
    Materials for Health Lab, Department of Chemical Engineering, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
  • James R. Runge
    James R. Runge
    University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.
    Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
  • Matthew Oshinowo
    Matthew Oshinowo
    University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.
    Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
  • Hannah S. Leese*
    Hannah S. Leese
    University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.
    Materials for Health Lab, Department of Chemical Engineering, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
    *Hannah Leese. Email: [email protected]
  • Antoine Buchard*
    Antoine Buchard
    University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.
    Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
    *Antoine Buchard. Email: [email protected]
Open PDFSupporting Information (1)

ACS Applied Energy Materials

Cite this: ACS Appl. Energy Mater. 2023, 6, 5, 2924–2935
Click to copy citationCitation copied!
https://doi.org/10.1021/acsaem.2c03937
Published February 21, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

This report describes the synthesis and characterization of organogels by reaction of a diol-containing polyether, derived from the sugar d-xylose, with 1,4-phenylenediboronic acid (PDBA). The cross-linked materials were analyzed by infrared spectroscopy (FT-IR), thermal gravimetric analysis (TGA), scanning electron microscopy (FE-SEM), and rheology. The rheological material properties could be tuned: gel or viscoelastic behavior depended on the concentration of polymer, and mechanical stiffness increased with the amount of PDBA cross-linker. Organogels demonstrated self-healing capabilities and recovered their storage and loss moduli instantaneously after application and subsequent strain release. Lithiated organogels were synthesized through incorporation of lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) into the cross-linked matrix. These lithium–borate polymer gels showed a high ionic conductivity value of up to 3.71 × 10–3 S cm–1 at 25 °C, high lithium transference numbers (t+ = 0.88–0.92), and electrochemical stability (4.51 V). The gels were compatible with lithium-metal electrodes, showing stable polarization profiles in plating/stripping tests. This system provides a promising platform for the production of self-healing gel polymer electrolytes (GPEs) derived from renewable feedstocks for battery applications.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2023 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

The reversible reaction of boronic acids with cis-1,2 or cis-1,3 diols to form boronate esters has been leveraged in applications such as glucose and chemosensors, (1−6) drug delivery systems, (7−9) and chromatography. (10−12) When combined with polymeric materials, this reversible chemistry can generate functional, cross-linked hydro- or organogel networks. The formation of boronate ester cross-links is reversible, depending on pH, temperature, and the presence of other hydroxylated compounds. (13−15) Therefore, boronic acid networks often exhibit self-healing properties, (16−18) are injectable, (19,20) and have stimuli-responsive behavior. (21) The alginate-boronic acid hydrogel developed by Hong et al. demonstrates for example many of the desirable properties of boronic acid-diol cross-linking: it shows self-healing and high strain to failure and is responsive to pH and glucose concentration. (22)
Many of these self-healing networks are made from reaction of a boronic-acid-containing polymer (often a polymer containing phenylboronic acid (PBA) side chains) with a second, diol-containing polymer, such as poly(vinyl alcohol) (PVA), or catechol derivatives of polymers. (14,16−18,23) Alternatively, both components can exist within the same polymer, leading to intramolecular cross-linking and gel formation. (20−22)
Cross-linking can also be achieved by reaction of a diol-containing polymer with a small molecule, a diboronic acid cross-linker. Cross-linking of PVA with borax or boric acid, for example, is well-known to easily form a gel. (24−26) 1,4-Phenylenediboronic acid (PDBA) is soluble in many organic solvents and so can be used to synthesize various organogels. Duncan et al. demonstrated the formation of dimethyl sulfoxide (DMSO), dimethylformamide (DMF), and methanol gels from PDBA and hydrolyzed poly(vinyl acetate) (PVAc). (27,28) Gelation between PDBA and PVA or guar is also possible, with the latter able to form hollow microspheres by the extrusion of PVA into PDBA solutions. (29,30) Hydrogels of PDBA can also be formed indirectly, by immersion of organogels in water. Nishiyabu et al. for example combined PDBA, PVA, and a dansyl-modified boronic acid in DMSO to afford a fluorescent gel capable of the detection of aqueous copper ions. (3) A hydrogel has also been made from the meta derivative, 1,3-phenylenediboronic acid, and catechol-derivatized poly(ethyleneglycol) (PEG). (13)
The Lewis acidity of boron allows further functionality to be installed into the networks to create ion-conductive organogels for application as gel polymer electrolytes (GPEs). GPEs are attractive materials for use in lithium-ion technologies as they show improved safety compared to liquid electrolytes while maintaining high ionic conductivity relative to inorganic and solid polymer electrolytes (SPEs). (31,32) Indeed, boron atoms possess vacant p-orbitals which can coordinate to Lewis bases, such as hydroxide anions. (12,33) This forms a new sp3, anionic, tetrahedral boron center, immobilized within the polymer structure creating a single-ion conducting effect. Therefore, only the cations can dissociate through the polymer matrix, resulting in increased ionic conductivity and transference numbers close to unity. The electron-deficient sp2 boron center can also coordinate anions of lithium salts, acting as an “anion-trap”, enhancing dissociation of cation–anion pairs to further improve the conductivity and transference number. (33,34)
Several groups have exploited the single-ion conducting (35−40) and anion-trapping (41−44) capabilities of boron in polymer electrolyte systems. Among these, Mecerreyes and co-workers have recently reported single-ion conducting GPEs with high ambient ionic conductivity (0.71 × 10–3 S cm–1 at 25 °C) and transference numbers of up to 0.85 based upon lithium borate methacrylate polymers. (45,46) Meanwhile, Chen et al. have developed a hydrogel electrolyte through the copolymerization of methacrylate and acrylamide monomers in the presence of borate. The resulting GPE showed excellent ambient ionic conductivity of 4.5 × 10–3 S cm–1 while being able to self-heal during to the dynamic nature of its borate-diol cross-links. (47) The GPE developed by Shim et al. shows a similarly high ionic conductivity of 4.2 × 10–3 S cm–1 at 30 °C. This GPE, which is based upon a semi-interpenetrating network containing poly(vinylidene fluoride) (PVDF) and cross-linkers containing ethylene oxide chains and anion-trapping boron groups, also shows a high transference number of 0.82 due to the introduction of the boron moieties. (31) However, these examples, and most examples of boronic acid hydro- or organogels, rely on polymers derived from petrochemicals.
Our group, (48−52) among many others, (53−56) has identified monosaccharides as a promising sustainable feedstock for polymer synthesis, owing to their low cost, low toxicity, and high abundance as well as their potential for pre- or postpolymerization functionalization, due to the presence of multiple hydroxy groups. (57−59) Herein, we report a novel organogel system from the cross-linking of a sugar-derived polyether with PDBA. The polyether, previously studied in our group, is derived from d-xylose and can be modified by postpolymerization to reveal cis-1,2-diols along the polymer backbone. (50) These hydroxy groups have been utilized to generate DMSO organogels with tunable rheological properties and self-healing capabilities. Cross-linking in the presence of a lithium salt exploited the Lewis acidity of the boron-based cross-links and produced conductive organogels, which have been further investigated as a single-ion conducting gel polymer electrolyte.

2. Results and Discussion

Click to copy section linkSection link copied!

2.1. Organogel Synthesis

A sugar-derived, diol-containing polymer was synthesized through postpolymerization modification of a polyether derived from d-xylose (Scheme 1). As previously reported, (50) an oxetane-type monomer, d-1, was synthezised in three steps from d-xylose. Anionic ring-opening polymerization (ROP) of d-1 with potassium tert-butoxide and 18-crown-6, in a monomer:KOtBu:18-crown-6 ratio of 100:1:1, yielded a regioregular polyether, poly(d-1), with Mn,SEC of 12,500 g mol–1 (ĐM = 1.23; vs poly(lactide) standards in THF). By acid hydrolysis, near quantitative deprotection of acetal groups was achieved (>95% deprotection), yielding DMSO and water-soluble polyethers, dp-poly(d-1), with Mn,SEC of 20,000 g mol–1 (ĐM = 1.21; vs poly(methyl methacrylate) standards in DMAc/LiBr).

Scheme 1

Scheme 1. Synthesis of Monomer d-1, Subsequent Anionic Ring-Opening Polymerization, and Post-Polymerization Acetal Deprotection to Yield Hydroxy-Polymer, dp-poly(d-1)
The newly revealed hydroxy groups in dp-poly(d-1) could then be reacted with difunctional boronic acids to form cross-linked organogels. 0.5 equiv of PDBA (with respect to the deprotected monomer unit, assuming quantitative acetal deprotection) was added to a 0.287 mol L–1 solution of dp-poly(d-1) in DMSO (Scheme 2). When stirred at room temperature (rt), the formation of a gel occurred within 5 min, as proved qualitatively via a vial inversion test (Figures 1 a–c). Reaction mixtures in the absence of either dp-poly(d-1) or PDBA remained liquid. By altering either the equivalents of PDBA or the concentration of polymer in solution, the conditions required for gelation were found (Figure 1d). The formation of a gel, as determined by the vial inversion test, is more likely in samples with high polymer concentrations and PDBA equivalents (SI, Figures S11–S14). In some cases, gel-like material was observed, but vial inversion was not successful. Hence, a partial gel event was defined.

Scheme 2

Scheme 2. Cross-Linking of dp-poly(d-1) and 1,4-Phenylenediboronic Acid (PDBA)

Figure 1

Figure 1. Vial inversion tests of (a) a 24 mg mL–1 solution of PDBA in DMSO, (b) a 0.287 mol L–1 solution of dp-poly(d-1), and (c) a reaction between dp-poly(d-1) and PDBA observed over several hours. (d) Phase diagram of the dp-poly(d-1)/PDBA system in DMSO at room temperature.

The equilibrium between boronate esters and diols/boronic acids is known to be pH dependent, with dissociation into free boronic acid and diols favored at pH values below the pKa of the boronic acid. (7,13,16,21,30,41) As expected, the pH reduction of the gel solutions below 2 by addition of hydrochloric acid disrupted cross-linking and caused dissolution of the polymer network, as qualitatively indicated by failure of the vial inversion test. Similarly, gels were not stable under basic conditions, with dissolution occurring above pH 13. Small amounts of gel could be reformed by addition of acid to the basic solution (Figure S15).
FT-IR analysis revealed several new chemical environments in the gels, when compared to dp-poly(d-1) and PDBA. Particularly, cross-linking was evidenced by a new signal at 696 cm–1 corresponding to new boronate ester bonds, and at new C–O stretching frequencies at 1205 cm–1 (Figure 2), the intensity of the latter increases with the degree of cross-linking (Figures S16–S18). However, the extent of cross-linking cannot be determined, as after gelation the materials were insoluble in all common solvents, hindering NMR analysis. Monosubstituted moieties, in which PDBA reacts with just one diol unit as shown in Scheme S1, may also be present within the organogels.

Figure 2

Figure 2. FT-IR spectra of dp-poly(d-1) (black), PDBA (red), and organogel (blue) made from dp-poly(d-1) (0.287 mol L–1) and 0.5 equiv of PDBA (with respect to monomer unit).

2.2. Rheology and Self-Healing Properties

Rheology was employed to gain further insight into the organogel materials. Frequency sweep experiments from 0 to 100 Hz showed most materials to be viscoelastic (Figure 3a). At lower frequencies, typically less than 13 Hz, the storage modulus (G′) is less than the loss modulus (G″). Beyond a cross-over frequency (ωc, 13.5–49.6 Hz), a gel-like region exists in which G′ is greater than G″. Looking at the 0.287 mol L–1/PDBA (0.25 equiv.) system in detail, G′ (793 Pa at 1 Hz) < G″ (2737 Pa at 1 Hz) was observed at rotating frequencies below 13.5 Hz, while G′ (9963 Pa at 20 Hz) > G″ (6648 Pa at 20 Hz) at frequencies greater than 13.5 Hz (Figure 3a, black trace). Such frequency-dependent moduli, with G′ > G″ only at higher frequencies, indicate the presence of dynamic cross-links. (16) At long time scales (low frequencies) the network can reorganize with viscous behavior, while at short time scales (high frequencies) the cross-linked network is more rigid, with elastic behavior. (42) The organogel stiffness depends on its composition, with increased PDBA content yielding increased storage and loss moduli across all frequencies due to higher degrees of cross-linking. Similar trends have been observed in other PDBA organogel systems. (27) The crossover frequency also increased as samples became more cross-linked, occurring at 13.5, 22.3, and 24.6 Hz for materials made with 0.25, 0.50, and 1.00 equiv of PDBA, respectively. The polymer concentration also affected the rheological behavior of the cross-linked materials. Those made with 0.144 mol L–1 polymer solutions exhibit true gel behavior, with G′ > G″ across all frequencies (Figure 3b). (22) This polymer concentration produced materials with consistent gel-like behavior, irrespective of the amount of PDBA (Figures S22–S24).

Figure 3

Figure 3. (a) Dynamic frequency sweep of organogels made from dp-poly(d-1) (0.287 mol L–1) and increasing PDBA equiv. (b) Dynamic frequency sweep of organogels made from dp-poly(d-1) (increasing concentrations) and PDBA (1 equiv).

The self-healing ability of the system was also investigated. An organogel made from a 0.144 mol L–1 solution of dp-poly(d-1) with 0.50 equiv of PDBA was chosen for this analysis, as it showed gel-like behavior across all frequencies and demonstrated relatively high G′ and G″ values. First, the critical strain required to break the cross-linking network was identified by increasing the applied strain on 1 Hz oscillatory measurements. After 5% strain, the material showed liquid behavior, with G″ > G′, indicating disruption of the gel network (Figure S26). Upon strain release (back to 0.007%), the cross-links were able to reform, and the original moduli values were immediately restored (Figure 4). The organogel showed a G′ of 8457 Pa and G″ of 352 Pa before the application of strain. After release, the material recovered its gel-like behavior (G′ > G″), exhibiting average moduli of 8452 and 622 Pa for G′ and G″, respectively.

Figure 4

Figure 4. Change in moduli during a strain ramp of an organogel (0.144 mol L–1 dp-poly(d-1), 0.50 equiv of PDBA) measured at a frequency of 1 Hz. The percentage strain was increased from 0.007% to 100%; then strain was released; and the moduli were measured at 0.007% strain.

The ability of the organogel to self-heal was further demonstrated in step-strain measurements performed at 1 Hz (Figure 5). The percentage strain was varied in sequential strain steps: 0% to 10% to 0% to 20% to 0% to 40% to 0% to 60% to 0% to 80% to 0% to 100% to 0%. Each time the strain was relaxed (up to 80% strain), the material fully recovered its original storage modulus. Even after release of the highest strains (80 and 100% strain), 84 and 74% of the original storage, modulus was recovered, with average values of 8640 and 7810 Pa, respectively. These results show that the material quickly recovers its structural properties after experiencing mechanical strain.

Figure 5

Figure 5. Gel fracture and self-healing in sequential strain steps. Top shows the measured storage moduli at each strain step, as described in the bottom plot (0% → 10% → 0% → 20% → 0% → 40% → 0% → 60% → 0% → 80% → 0% → 100% → 0%, measured at 1 Hz).

2.3. Lyophilized Gel Microstructure and Thermal Properties

To investigate the microstructure of the gel materials, samples made from a 0.144 mol L–1 solution of dp-poly(d-1) with 0.25–1.00 equiv of PDBA were analyzed by field emission scanning electron microscopy (FE-SEM). To remove DMSO from the system, the gels were first flash frozen in liquid nitrogen and then dried under vacuum in a process akin to freeze-drying. Materials exhibited large, irregular pores throughout. The microstructure showed a dependency on the cross-linking density (Figure 6). Cross-sectional imaging of the 0.25 material showed many globular pores, of up to 0.124 mm in length (in an observed area of 0.474 mm2). Increasing the proportion of PDBA to 0.50 equiv formed a less dense structure with directional, elongated pores (up to 0.415 mm in length, observed from 4.226 mm2). Further increase of cross-linking density produced a denser structure, with fewer observable pores (up to 0.118 mm observed in 0.473 mm2, also see Figures S35–S38). This structure could be due to more prominent monosubstitution of polymer chains (see Scheme S1).

Figure 6

Figure 6. FE-SEM imaging of lyophilized organogels made from 0.144 M solutions of dp-poly(d-1) in DMSO with 0.25–1.00 equiv of PDBA.

The thermal characteristics of the lyophilized gels were analyzed. Compared to the non-cross-linked dp-poly(d-1), thermal gravimetric analysis (TGA) of dried gel samples showed that cross-linking increased thermal stability. The sample with 0.25 equiv of PDBA showed a Td,max of 368 °C, compared to 272 °C for dp-poly(d-1). This increased thermal stability is expected from the introduction of cross-linking and the presence of rigid phenyl rings. Increasing the amount of PDBA to 0.50 equiv provides more cross-linking opportunities, which is reflected by a further increase in the onset of thermal degradation to Td,max = 376 °C. As the amount of PDBA was increased further to 1.00 equiv, the onset of degradation decreased slightly, to 356 °C (Figure 7a). This change could again be explained by an excess of PDBA favoring monosubstitution of dp-poly(d-1), leaving more diol moieties along the polymer chains. The presence of diol moieties is known to reduce the thermal stability of this polyether (Td,max,poly(d-1) = 372 °C vs Td,max,dp-poly(d-1) = 272 °C, Figures S39–S40); (50) hence, monosubstitution would decrease the observed Td,max. (54) Most cross-linked materials showed an absence of glass transition, crystallization, and melting events by differential scanning calorimetry (DSC, Figures S49–S51), although a small glass transition can be observed in the material made with 0.50 equiv of PDBA (Tg = 112 °C, Figure S47).

Figure 7

Figure 7. TGA of (a) lyophilized and (b) nonlyophilized organogel samples made from dp-poly(d-1) and PDBA with a polymer concentration of 0.144 mol L–1 and PDBA equiv of 0.25–1.00. Annotations of the Td,max values of each sample.

Nonlyophilized materials were also analyzed by TGA. A similar trend to the lyophilized materials was observed: the thermal stability increased as the amount of cross-linking increased until monosubstitution became favored at high PDBA equiv (Figure 7b). TGA analysis of nonlyophilized materials also allowed the gel solvent content to be quantified. Samples were held at 100 °C for two hours to remove all the DMSO solvent (Figures S44–S46), resulting in 75, 68, and 71% mass loss for materials (and therefore mass solvent percentage) made with 0.50, 0.50, and 1.00 equiv of PDBA, respectively.

2.4. Gel Polymer Electrolytes and Electrochemical Properties

As well as facilitating gelation and self-healing, cross-linking with boronic acids allows additional functionality to be installed via the empty p-orbital of boron. After reacting with the diols of dp-poly(d-1), the boron center is sp2 hybridized and can coordinate with an additional nucleophilic group to become anionic and sp3 hybridized. (31) Balancing the charge with lithium counter cations introduces conductivity. Thus, cross-linking dp-poly(d-1) with PDBA (0.5 equiv) in the presence of lithium bis(trifluoromethanesulfonyl)imide (LiTFSI, 1.0 equiv) produced single-ion conductive organogels (Scheme 3).

Scheme 3

Scheme 3. Cross-Linking of dp-poly(d-1) and PDBA in the Presence of a Lithium Salt (LiTFSI)
The presence of LiTFSI in the cross-linked network changed the material’s rheological behavior. The material made from 0.144 mol L–1 solution of dp-poly(d-1) and 0.50 equiv of PDBA still showed gel behavior (G′ > G″) and self-healing capabilities (Figures S27–S29), akin to its nonlithiated analogue, but was softer with reduced storage and loss moduli (at 1 Hz, GLi = 2227 Pa vs Gnon,Li = 10233 Pa, GLi = 1099 Pa vs Gnon,Li = 2923 Pa). Materials made with a 0.287 mol L–1 solution of dp-poly(d-1) and 0.50 equiv of PDBA were viscoelastic in the absence of LiTFSI but showed gel-like behavior and self-healing when lithiated (Figures S30–S32). The effect LiTFSI had on the rheological properties of these materials confirms that it is fully incorporated within the cross-linked polymeric network.
The temperature within electrochemical devices is variable during battery operation. Therefore, temperature-dependent rheological measurements were conducted on the lithiated gels, to gain further insight into their mechanical integrity. Both materials retained gel-like behavior across the temperature range studied (25–70 °C). Although the 0.144 mol L–1 material softened with increasing temperature (decrease of storage and loss moduli, Figure S33), the 0.287 mol L–1 material showed good thermal stability, with only small differences between the storage and loss moduli measured at 25 and 70 °C (G25 °C = 1237 Pa, G75 °C = 1037 Pa, 84% decrease, Figure S34). Retention of properties at elevated temperatures indicated that this latter material was suitable for electrochemical analysis and had potential to be a self-healing, renewable gel polymer electrolyte.
The conductivity of this organogel (referred to as Li-gel in the following) was assessed using electrochemical impedance spectroscopy (EIS). Typical Nyquist plots were obtained, with partial semicircles representing bulk resistance (Rb) followed by diagonal lines showing Warburg diffusion (Figure S52). The gel showed a room temperature (25 °C) ionic conductivity of 3.71 × 10–3 S cm–1. As shown by the Arrhenius plots in Figure 8, conductivity was further enhanced at elevated temperatures as Rb decreased, reaching 4.99 × 10–3 S cm–1 at 60 °C (Figure S54, Table S1). As a control, a nonlithiated gel analogue was also analyzed by EIS and showed reduced ionic conductivity across all temperatures compared to the sample containing LiTFSI (Figures S53–S54, Table S2). These initial EIS experiments demonstrate the potential for this system to be used as a gel polymer electrolyte, or GPE. GPEs are attractive electrolyte materials as they alleviate the safety concerns of liquid electrolytes while achieving higher ionic conductivities than solid polymer electrolyte (SPE) alternatives. (31,32) The ambient ionic conductivity of our system (3.71 × 10–3 S cm–1 at 25 °C) matches or exceeds recently reported, boron-containing GPEs (see Table S3). (60−63)

Figure 8

Figure 8. Temperature dependence of ionic conductivity (σ) for lithiated (1 equiv of LiTFSI, 0.144 mol L–1; Li-gel) and nonlithiated (control) organogels made from dp-poly(d-1) (0.287 mol L–1 with respect to the repeating unit) with 0.50 equiv of PDBA.

In terms of single-ion conducting GPEs featuring anion-trapping boron moieties, the ionic conductivity values reported for Li-gel in this study are comparable to those achieved by Ma and co-workers (2.33 × 10–3 S cm–1 at room temperature) (60) and Dai et al. (8.4 × 10–4 S cm–1 at 30 °C). (61) Similarly, Shi and co-workers have recently reported a novel single-ion conducting boron-centered GPE, which showed ionic conductivity (1.03 × 10–3 S cm–1 at 32 °C) in the same order of magnitude as the system reported in this study. (62) Moreover, Xu et al. have reported a self-healing hydrogel polymer electrolyte, based upon copolymers of glycerol monomethacrylate and acrylamide with dynamic borate cross-links, which showed comparable ionic conductivity of 4.5 × 10–3 S cm–1 at room temperature. (63)
Further electrochemical characterization was carried out. Electrochemical stability was determined by linear sweep voltammetry (Figure S56). Li-gel displayed anodic stability up to 4.51 V (vs Li/Li+), therefore beyond the 4.2 V stability to oxidation required for practical application. The lithium transference number was also determined using the Bruce–Vincent method, combining DC polarization chronoamperometry and EIS measurements in a symmetrical Li|GPE|Li cell. Although equilibrium was not obtained, Li-gel presented high tLi+ values between 0.88 and 0.92 during combined EIS and chronoamperometry measurements (average of 3 repeats, Table S4 and Figure S57). This data provides an estimate of the tLi+ of Li-gel and is consistent with single-ion conducting behavior, demonstrating the potential of the GPE.
Finally, the performance of Li-gel for the stripping/plating behavior of Li+ was investigated by recording polarization profiles in a symmetrical Li|GPE|Li cell at 25 °C (Figure 9). Li-gel exhibited good cycling behavior for at least 10 full cycles (20 h) at consecutive current densities of 0.1, 0.5, and 1.0 mA cm–2, with small overpotentials ≤0.05 V (vs Li/Li+) for the latter current density. When a large current density of 2.5 mA cm–2 was applied, a severe voltage noise was observed which was attributed to short circuits due to lithium dendrites emerging from the electrode surfaces. (64) Nevertheless, the ability of the material to withstand a critical current density of up to 1.0 mA cm–2 is highly promising and compares well with recent single-ion gel polymer electrolytes based on borate pendant groups, even if more studies are needed, including over a longer time frame and in Li|Li-gel|cathode cells. However, preliminary results using Li|Li-gel|LiFePO4 showed poor charging/discharging behavior, suggesting that lithium iron phosphate may not be an appropriate cathode material for this GPE. The dynamic nature of boronic acid cross-linking and the generation of water upon cross-linking may also explain this and the fluctuation in current during the measurement of tLi+. More investigations are currently underway and will be reported in due course.

Figure 9

Figure 9. Li+ stripping/plating voltage profile of Li-gel in a symmetrical Li|Li-gel|Li cell at 25 °C. The half-cycle duration was 1 h with a cutoff voltage of ±1.0 V.

Like the majority of polymer electrolytes, the leading boron-containing GPEs rely on fossil fuel derived polymers. Several GPEs have recently been developed from renewable, biobased materials, such as derivatives of cellulose, natural rubber, and chitosan. (65−69) The electrochemical performance of the sugar-derived system reported here rivals many of these examples and combines the benefits of anion trapping with the sustainability benefits of renewable polymers. Further investigation is therefore warranted to study and optimize the anion-trapping capabilities of PDBA in these systems and its effect on the electrochemical performance of the renewable and self-healing GPE.

3. Conclusions

Click to copy section linkSection link copied!

A new self-healing organogel system has been developed from the room-temperature cross-linking of a sugar-derived polyether and 1,4-phenylenediboronic acid (PDBA) in DMSO. The effect of cross-linking density and polymer concentration on the material properties has been demonstrated, with viscoelastic and gel materials favored at polymer concentrations above and below 0.144 mol L–1, respectively. As expected, the stiffness of the materials can be tuned according to the equivalents of PDBA: stiffer materials are produced from higher cross-linking densities. Due to the dynamic nature of the boronate ester bonds, the materials show excellent self-healing abilities. After application of up to 100% strain, materials immediately recover at least 74% of their original storage modulus. The boronate ester moieties also allowed lithium ions to be incorporated throughout the cross-linked polymer matrix network, producing single-ion conductive organogels. Gels retain their self-healing nature after lithiation, enabling the production of self-healing, gel polymer electrolytes (GPEs). Initial electrochemical impedance spectroscopy (EIS) measurements show that this system is a promising candidate for Li battery applications. The organogel demonstrates high ambient ionic conductivities of 3.71 × 10–3 S cm–1, matching or outperforming recently reported GPEs. It also showed high anodic stability (4.51 V vs Li/Li+), high transference numbers (0.88–92), and high critical current density (1 mA cm–2). This promising electrical performance, coupled with the self-healing properties and renewable nature of the xylose-derived polymer, means there is scope to optimize electrochemical performance (including in whole cell setups) while investigating gelation in common electrolyte solutions, such as ethylene (EC), propylene (PC), or dimethyl carbonate (DMC).

4. Experimental Section

Click to copy section linkSection link copied!

4.1. Materials and Methods

All reagents were used all purchased, without further purification. d-Xylose and 1,2-O-isopropylidene-α,d-xylofuranose were purchased from Carbosynth. Sulfuric acid (H2SO4) was purchased from VWR. Potassium carbonate (K2CO3), trifluoroacetic acid (TFA), triethylamine (NEt3), and potassium methoxide (KOMe) were purchased from Fischer scientific. Tosylchloride (TsCl), potassium tert-butoxide (KOtBu), and 1,4-phenylenediboronic acid (PDBA) were purchased from Merck. 18-Crown-6 was purchased from Acros Organics.

4.2. Characterization Methods

4.2.1. NMR Spectroscopy

NMR spectra were recorded in either CDCl3, DMSO, or D2O on a Bruker-400 or -500 MHz instrument referenced to the corresponding residual solvent as an internal standard: 1H NMR spectra (400 or 500 MHz) δH = 7.26, δH = 2.50, and δH 4.79 ppm for residual protiated CDCl3, DMSO, and D2O, respectively: 13C NMR spectra (101 or 126 MHz) δC = 77.16, δC = 39.52 ppm for CDCl3, DMSO-d6, and D2O, respectively.

4.2.2. Size-Exclusion Chromatography (SEC)

The number-average molar mass, Mn, and dispersity, ĐM, of polymer samples were determined by size exclusion chromatography (SEC) in a THF or dimethylacetamide (DMAc) eluent (DMAc contains 0.1% w/v LiBr). Polymer samples were dissolved to a concentration of ∼2 mg/mL. Multianalysis software was used to process the data. THF samples were recorded on an Agilent 1260 Infinity series instrument at 1 mL min–1 at 35 °C using two PLgel 5 μm MIXED-D 300 × 7.5 mm columns in series. Samples were detected with a differential refractive index (RI) detector. Number-average molar mass (Mn,SEC) and dispersities (ĐM (Mw/Mn)) were calculated against a polystyrene calibration (11 polystyrene standards of narrow molar mass, ranging from Mw 615 to 568000 Da). DMAc samples were measured on an Agilent 1260 SEC MDS instrument at 0.5 mL min–1 at 50 °C using a Polargel-M 300 × 7.5 mm column. Samples were detected using an RI detector. Mn,SEC and ĐM were calculated against a polymethylmethacrylate calibration (11 polymethylmethacrylate standards of narrow molar mass ranging from Mw 885 to 260900).

4.2.3. Thermogravimetric Analysis (TGA)

TGA was performed on a Setsys Evolution TGA 16/18 from Setaram. The Calisto program was used to collect and process data. Samples were loaded into a 170 μL alumina crucible and heated from 30 to 500 °C at a rate of 10 °C min–1.

4.2.4. Differential Scanning Calorimetry (DSC)

Glass transition temperatures, Tg, of polymer samples were found using DSC on a Q20 machine from TA Instruments, controlled by the Q Series program. Samples were loaded into a 10 μL Tzero aluminum pan. Samples were heated and cooled at a rate of 10 °C min–1 and immediately submitted to a second heating and cooling cycle at the same rate. The heat flow was recorded against an empty reference 10 μL Tzero aluminum reference pan.

4.2.5. IR Spectroscopy

IR spectra were recorded using a Spectrum 100 FT-IR spectrometer from Perkin-Elmer and were baseline corrected in MatLab.

4.2.6. pH

Measurements were conducted using an MD 8000 L pHenomenal pH meter from VWR.

4.2.7. Rheology

Measurements were conducted using a Discovery HR-2 hybrid rheometer from TA Instruments with a 25 mm parallel plate geometry. Unless otherwise specified, a temperature of 25 °C was maintained throughout the experiments. Oscillatory frequency sweep measurements were conducted at 0.007% strain. Oscillatory strain amplitude measurements were conducted at a frequency of 1 Hz.

4.2.8. Microscopy

Images were taken using a JEOL JSM-7900F field emission scanning electron microscope (FE-SEM) from JEOL U.K., Ltd. Before imaging, samples were coated with 20 nm of gold using a Quorum Q150T S sputter coater from Quorum Tech. Analysis of images was conducted using the Image J software (Java 1.8.0_112 64 bit).

4.2.9. Electrochemistry

Measurements were performed using a modified version of a TCS battery cell (RHD instruments) with blocking stainless steel current collectors (exposed area diameter = 0.8 cm) connected to a Metrohm Autolab PGSTAT204 potentiostat with a FRA32M module. The cell components were dried in a vacuum oven at 70 °C prior to assembly. Temperature control of the cell was achieved using an Autolab Microcell HC temperature-controlled cell stand and temperature controller designed by RHD instruments. The cell was equilibrated for 1 h at each temperature before measurements were taken. Ionic conductivity (σ) was determined by a two-electrode electrochemical impedance spectroscopy (EIS) measurement in the typical frequency range of 0.1 Hz to 0.5 MHz with an applied amplitude of 50 mV in a symmetrical SS|GPE|SS cell. Results shown are averages of five repeat measurements taken at each temperature value. NOVA 2.1 (Metrohm) software was used to analyze the results. The bulk resistance (Rb) was taken as the high-frequency intercept of the Z′ axis on the Nyquist plot. (70) The ionic conductivity values were then calculated using
σ=lRbA
where l is the thickness of the gel (determined by triplicate measurements with digital callipers), and A is the area of the electrode surface.
Linear sweep voltammetry was used to determine the electrochemical stability in a Li|GPE|SS cell using a lithium counter/reference electrode at 25 °C. The open-cell voltage (OCV) was first determined, and then the voltage swept from the OCV to +6 V with a scan rate of 1 mV s–1.
Lithium transference number (t+) was determined using a combined EIS and chronoamperometry method using a Li|GPE|Li symmetrical cell using lithium foil of 0.75 mm thickness and an applied voltage of 10 mV. The measurement was recorded at 25 °C, and t+ was calculated using the Bruce–Vincent equation
t+=Iss(ΔVI0Rb,0)I0(ΔVIssRb,SS)
where ΔV is the applied voltage and I0 and ISS represent the initial and steady state current before and after DC polarization. Rb,0 and Rb,SS represent the bulk resistance obtained from EIS measurements before and after DC polarization.
The Li+ stripping/plating behavior was measured in symmetrical Li/Li cells (Li thickness = 750 μm) with half-cycle durations of 1 h at 25 °C. Ten full cycles with ±1.0 V cutoff voltages were performed for each current density until the experiment was stopped due to cell failure.

4.3. Synthetic Procedures

4.3.1. 1,2-O-Isopropylidene-xylofuranose (IPXF)

Following an adapted literature procedure: (1)d-xylose (20.00 g, 133.34 mmol, 1.00 equiv), nonanhydrous acetone (500 mL), and 95–98% concentrated H2SO4 (20 mL, 373.16 mmol, 2.80 equiv) were combined to form a pale-yellow suspension. The suspension was left to stir until the solid had fully dissolved to form a yellow solution (ca. 1 h). Upon full dissolution, an aqueous solution of K2CO3 (224 mL, 1.10 mol L–1, 246.01 mmol, 1.84 equiv) was slowly added. The resultant white suspension was allowed to stir at room temperature with aliquots taken at regular intervals to monitor the deprotection by TLC (1:1 HCCl3:acetone eluent) and NMR (d6-DMSO). After approximately 90 min, near-quantitative formation of the monoprotected sugar was observed. Solid K2CO3 (17.80 g, 128.79 mmol, 1.04 equiv) was added slowly at room temperature and the pH adjusted until 7–8 by litmus paper. The suspension was then filtered and the acetone removed in vacuo at 40 °C. The aqueous phase was washed with DCM (100 mL, ×3), and the organic phases were collected and back extracted with water (50 mL, ×3). The aqueous phases were collected, and the water was removed in vacuo at 50 °C. The resultant oil was dissolved in EtOAc and stirred over MgSO4 overnight. The suspension was filtered and the solvent removed in vacuo at 40 °C to yield 1,2-O-isopropylidene-α-xylofuranose (IPXF) as a clear oil (19.76 g, 78%).

4.3.2. 1,2-O-Isopropylidene-5-O-tosyl-xylofuranose (Ts-IPXF)

IPXF (20.00 g, 105.16 mmol, 1.00 equiv) and tosyl chloride (22.06 g, 115.68 mmol, 1.10 equiv) were dissolved in dichloromethane (200 mL). The reaction vessel was cooled over ice, and triethylamine (100 mL, 574.17 mmol, 5.46 equiv) was added. The vessel was left to warm to 20 °C overnight. Ethyl acetate (200 mL) was added, and the mixture was transferred to a separating funnel. The organic phase was then washed with brine (150 mL, ×1), 1 mol L–1 sodium hydrogen carbonate (150 mL, ×1), and water (150 mL, ×1). The organic phase was collected and dried over magnesium sulfate, filtered, and concentrated in vacuo at 40 °C to give a white solid. The solid was then stirred over cold Et2O (−18 °C, 150 mL) for 10 min, filtered over a glass frit, and rinsed again with cold Et2O (50 mL). The precipitate was left to dry in a vacuum oven overnight at 50 °C to give the product as an off-white solid (28.32 g, 82.59 mmol, 79%). Synthesized or commercial IPXF can be used. Quoted yield uses commercial IPXF.

4.3.3. Monomer d-1

To a round-bottom flask was charged Ts-IPXF (15.90 g, 46.04 mmol, 1.00 equiv) and acetonitrile (160 mL). The suspension was agitated at 400 rpm until a clear colorless homogeneous solution was obtained. Potassium methoxide (6.84 g, 96.68 mmol, 2.10 equiv) was added to the solution which was then heated to reflux (80 °C) for 1 h. The resulting dark brown reaction was quenched by addition of water (80 mL). The acetonitrile was removed in vacuo at 40 °C. The aqueous phase was extracted with Et2O (100 mL, ×3). The organic phases were collected and washed with brine (100 mL, ×1), 1 mol L–1 sodium hydrogen carbonate (100 mL, ×1), and water (100 mL, ×1). The organic phases were collected, dried over MgSO4, filtered, and concentrated in vacuo at 40 °C to give d-1 as a clear oil. For further purification, the oil was stirred over CaH2 at 80 °C overnight and vacuum distilled (1 × 10–2 mbar, 40 °C) to yield the oxetane as a clear oil (5.52 g, 32.11 mmol, 70%).

4.3.4. General Procedure for the Polymerization of d-1

Under an argon atmosphere, a centrifuge tube was charged with the oxetane monomer (2.00 g, 11.63 mmol, 100.00 equiv), KOtBu (232 μL, 0.5 mol L–1 in THF, 1.00 equiv), and 18-crown-6 (232 μL, 0.5 mol L–1 in THF, 1.00 equiv). The tube was sealed and heated to 120 °C with stirring. After 22 h, the vial was cooled, and the solid was dissolved in the minimum amount of CHCl3 and then precipitated from cold Et2O. The suspension was centrifuged (3500 rpm, 5 min), and the solid phase was collected. The polymer was then redissolved in CHCl3 and precipitated twice more from cold Et2O with centrifugation (3500 rpm, 5 min). The solid was collected and dried in a vacuum oven for 24 h at 100 °C to yield the polyether (1.59 g, 9.26 mmol, 80%).

4.3.5. General Procedure for the Deprotection of Polyethers

Following an adapted literature procedure: to a dram vial was added poly(d-1) (1.86 g, 10.83 mmol, Mn,SEC 7250–12,600 g mol–1, ĐM = 1.2) and DCM (7.40 mL). Upon dissolution, the reaction mixture was cooled to 0 °C, and a 4:1 TFA:H2O solution (18.00 mL) was added. After 8 h, the produce was precipitated from cold Et2O, and the resulting suspension was centrifuged (3500 rpm, 5 min). The solid phase was collected and rinsed twice more in cold Et2O or until the supernatant was neutral by a litmus test. The solid phase was collected, but not dried, to yield the polyether deprotected at 91% (calculated by relative integration of protected and deprotected anomeric environments in 1H NMR spectroscopy, see Section 4.2.5). The product was immediately dissolved into an appropriate solvent for analysis or further reaction (1.40 g, 10.60 mmol, 98%, Mn,SEC 20,300–24,300 g mol–1, ĐM = 1.21–1.71).

4.3.6. General Procedure for the Cross-Linking of dp-poly(d-1) and PDBA

In a dram vial, PDBA (0.05–1.00 equiv with respect to the deprotected polymer repeat unit; an example calculation of the amount of PDBA used in the synthesis can be found in the Supporting Information) was added to a solution of dp-poly(d-1) in DMSO (0.072–0.409 mol L–1). The mixture was stirred at room temperature until gelation occurred (as confirmed by a vial inversion test). For rheological and self-healing measurements, the mixture was transferred into a glass Petri dish (diameter = 3.5 cm, depth = approximately 0.8 cm) before the point of gelation. The mixture was covered and allowed to mature overnight at room temperature.

4.3.7. General Procedure for the Formation of Conductive Organogels

In a dram vial, PDBA (0.50 equiv with respect to the deprotected polymer repeat unit) and lithium bis(trifluoromethanesulfonyl)imide (LiTFSI, 1.00 equiv) were added to a solution of dp-poly(d-1) in DMSO (0.144 or 0.287 mol L–1). The mixture was stirred at room temperature and then was transferred into a glass Petri dish (diameter = 3.5 cm, depth = approximately 0.8 cm) before the point of gelation. The mixture was covered and allowed to mature at room temperature. To perform EIS, a disc of diameter 1.1 cm was cut from the gel after 3–4 h of curing. The gel was then transferred into the electrochemical cell for analysis: the disc of gel was placed onto a stainless-steel current collector (diameter of exposed area = 0.8 cm) electrode inside a TSC battery cell (RHD Instruments). The second electrode was placed on top of the gel, and the cell was closed and tightened to ensure sufficient electrode contact. For LSV and DC polarization measurements where lithium electrodes (d = 1.0 cm) were used, the electrodes were first polished with a nylon brush before contact with the gel was made.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsaem.2c03937.

  • Cross-linking procedures, NMR spectra, photographs of vial inversion tests, FT-IR spectra, detailed rheological characterizations, FE-SEM microscopy, TGA and DSC traces, and additional electrochemistry results (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Emma L. Daniels - University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.Materials for Health Lab, Department of Chemical Engineering, University of Bath, Claverton Down, Bath BA2 7AY, U.K.Orcidhttps://orcid.org/0000-0002-3227-049X
    • James R. Runge - University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
    • Matthew Oshinowo - University of Bath Institute for Sustainability, Claverton Down, Bath BA2 7AY, U.K.Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. ELD: polymer and material synthesis and characterization, investigation, and manuscript writing. JRR and MO: electrochemical analysis, methodology, and manuscript writing. Electrochemical analysis was performed by JRR. HSL and AB: conceptualization, manuscript writing – review and editing, supervision, and funding acquisition.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank Konstantinos Myronidis for help and useful discussions regarding rheological measurements and Prof Frank Marken for advice on EIS. Analytical facilities were provided through the Material and Chemical Characterization Facility (MC2) at the University of Bath. We thank Dr. Philip Fletcher for assistance with FE-SEM imaging. Research funding from the Engineering and Physical Sciences Research Council (DTP studentship to JRR, EP/L016354/1 CDT in Sustainable Chemical Technologies Studentship to ELD), the University of Bath (studentship to MO), and the Royal Society (UF/160021 and URF\R\221027, fellowship to AB) is also acknowledged.

ABBREVIATIONS

Click to copy section linkSection link copied!

PDBA

1,4-phenylenediboronic acid

FT-IR

Fourier-transform infrared

TGA

thermal gravimetric analysis

DSC

differential scanning calorimetry

LiTFSI

lithium bis(trifluoromethanesulfonyl)imide

GPE

polymer gel electrolyte

PVA

poly(vinyl alcohol)

DMSO

dimethyl sulfoxide

DMF

dimethylformamide

PVAc

poly(vinyl acetate)

PEG

poly(ethyleneglycol)

SPE

solid polymer electrolyte

ROP

ring-opening polymerization

DMAc

dimethylacetamide

LiBr

lithium bromide

EIS

electrochemical impedance spectroscopy

DC

direct current

SEC

size exclusion chromatography

THF

tetrahydrofuran

References

Click to copy section linkSection link copied!

This article references 70 other publications.

  1. 1
    Asher, S. A.; Alexeev, V. L.; Goponenko, A. V.; Sharma, A. C.; Lednev, I. K.; Wilcox, C. S.; Finegold, D. N. Photonic Crystal Carbohydrate Sensors: Low Ionic Strength Sugar Sensing. J. Am. Chem. Soc. 2003, 125 (11), 33223329,  DOI: 10.1021/ja021037h
  2. 2
    Lerner, M. B.; Kybert, N.; Mendoza, R.; Villechenon, R.; Bonilla Lopez, M. A.; Charlie Johnson, A. T. Scalable, non-invasive glucose sensor based on boronic acid functionalized carbon nanotube transistors. Appl. Phys. Lett. 2013, 102 (18), 183113,  DOI: 10.1063/1.4804438
  3. 3
    Nishiyabu, R.; Kobayashi, H.; Kubo, Y. Dansyl-containing boronate hydrogel film as fluorescent chemosensor of copper ions in water. RSC Adv. 2012, 2 (16), 65556561,  DOI: 10.1039/c2ra20516e
  4. 4
    Awino, J. K.; Gunasekara, R. W.; Zhao, Y. Selective Recognition of d-Aldohexoses in Water by Boronic Acid-Functionalized, Molecularly Imprinted Cross-Linked Micelles. J. Am. Chem. Soc. 2016, 138 (31), 97599762,  DOI: 10.1021/jacs.6b04613
  5. 5
    Chen, W.; Guo, Z.; Yu, H.; Liu, Q.; Fu, M. Molecularly imprinted colloidal array with multi-boronic acid sites for glycoprotein detection under neutral pH. J. Colloid Interface Sci. 2022, 607, 11631172,  DOI: 10.1016/j.jcis.2021.09.048
  6. 6
    He, P.; Zhu, H.; Ma, Y.; Liu, N.; Niu, X.; Wei, M.; Pan, J. Rational design and fabrication of surface molecularly imprinted polymers based on multi-boronic acid sites for selective capture glycoproteins. Chem. Eng. J. 2019, 367, 5563,  DOI: 10.1016/j.cej.2019.02.140
  7. 7
    Zhao, Z. W.; Yao, X. M.; Zhang, Z.; Chen, L.; He, C. L.; Chen, X. S. Boronic Acid Shell-Crosslinked Dextran-b-PLA Micelles for Acid-Responsive Drug Delivery. Macromol. Biosci. 2014, 14 (11), 16091618,  DOI: 10.1002/mabi.201400251
  8. 8
    Li, Y. P.; Xiao, W. W.; Xiao, K.; Berti, L.; Luo, J. T.; Tseng, H. P.; Fung, G.; Lam, K. S. Well-Defined, Reversible Boronate Crosslinked Nanocarriers for Targeted Drug Delivery in Response to Acidic pH Values and cis-Diols. Angew. Chem., Int. Ed. 2012, 51 (12), 28642869,  DOI: 10.1002/anie.201107144
  9. 9
    Su, J.; Chen, F.; Cryns, V. L.; Messersmith, P. B. Catechol Polymers for pH-Responsive, Targeted Drug Delivery to Cancer Cells. J. Am. Chem. Soc. 2011, 133 (31), 1185011853,  DOI: 10.1021/ja203077x
  10. 10
    Jiang, H.-P.; Qi, C.-B.; Chu, J.-M.; Yuan, B.-F.; Feng, Y.-Q. Profiling of cis-Diol-containing Nucleosides and Ribosylated Metabolites by Boronate-affinity Organic-silica Hybrid Monolithic Capillary Liquid Chromatography/Mass Spectrometry. Sci. Rep. 2015, 5 (1), 7785,  DOI: 10.1038/srep07785
  11. 11
    Koyama, T.; Terauchi, K.-i. Synthesis and application of boronic acid-immobilized porous polymer particles: a novel packing for high-performance liquid affinity chromatography. J. Chromatogr. B: Biomed. Sci. Appl. 1996, 679 (1), 3140,  DOI: 10.1016/0378-4347(96)00006-0
  12. 12
    Brooks, W. L. A.; Sumerlin, B. S. Synthesis and Applications of Boronic Acid-Containing Polymers: From Materials to Medicine. Chem. Rev. 2016, 116 (3), 13751397,  DOI: 10.1021/acs.chemrev.5b00300
  13. 13
    He, L. H.; Fullenkamp, D. E.; Rivera, J. G.; Messersmith, P. B. pH responsive self-healing hydrogels formed by boronate-catechol complexation. Chem. Commun. 2011, 47 (26), 74977499,  DOI: 10.1039/c1cc11928a
  14. 14
    Figueiredo, T.; Cosenza, V.; Ogawa, Y.; Jeacomine, I.; Vallet, A.; Ortega, S.; Michel, R.; Olsson, J. D. M.; Gerfaud, T.; Boiteau, J. G.; Jing, J.; Harris, C.; Auzely-Velty, R. Boronic acid and diol-containing polymers: how to choose the correct couple to form ″strong″ hydrogels at physiological pH. Soft Matter 2020, 16 (15), 36283641,  DOI: 10.1039/D0SM00178C
  15. 15
    Brooks, W. L. A.; Deng, C. C.; Sumerlin, B. S. Structure-Reactivity Relationships in Boronic Acid-Diol Complexation. ACS Omega 2018, 3 (12), 1786317870,  DOI: 10.1021/acsomega.8b02999
  16. 16
    Deng, C. C.; Brooks, W. L. A.; Abboud, K. A.; Sumerlin, B. S. Boronic Acid-Based Hydrogels Undergo Self-Healing at Neutral and Acidic pH. ACS Macro Lett. 2015, 4 (2), 220224,  DOI: 10.1021/acsmacrolett.5b00018
  17. 17
    Smithmyer, M. E.; Deng, C. C.; Cassel, S. E.; LeValley, P. J.; Sumerlin, B. S.; Kloxin, A. M. Self-Healing Boronic Acid-Based Hydrogels for 3D Co-cultures. ACS Macro Lett. 2018, 7 (9), 11051110,  DOI: 10.1021/acsmacrolett.8b00462
  18. 18
    Wang, X.-T.; Deng, X.; Zhang, T.-D.; Zhang, X.; Shi, W.-P.; Lai, J.; Zhou, H.; Ye, Y.-J.; Zhang, C.-Y.; Yin, D.-C. Biocompatible self-healing hydrogels based on boronic acid-functionalized polymer and laponite nanocomposite for water pollutant removal. Environ. Chem. Lett. 2022, 20 (1), 8190,  DOI: 10.1007/s10311-021-01350-4
  19. 19
    Huang, Z. J.; Delparastan, P.; Burch, P.; Cheng, J.; Cao, Y.; Messersmith, P. B. Injectable dynamic covalent hydrogels of boronic acid polymers cross-linked by bioactive plant-derived polyphenols. Biomater. Sci. 2018, 6 (9), 24872495,  DOI: 10.1039/C8BM00453F
  20. 20
    Dong, Y. Z.; Wang, W. H.; Veiseh, O.; Appel, E. A.; Xue, K.; Webber, M. J.; Tang, B. C.; Yang, X. W.; Weir, G. C.; Langer, R.; Anderson, D. G. Injectable and Glucose-Responsive Hydrogels Based on Boronic Acid-Glucose Complexation. Langmuir 2016, 32 (34), 87438747,  DOI: 10.1021/acs.langmuir.5b04755
  21. 21
    Pettignano, A.; Grijalvo, S.; Haring, M.; Eritja, R.; Tanchoux, N.; Quignard, F.; Diaz, D. D. Boronic acid-modified alginate enables direct formation of injectable, self-healing and multistimuli-responsive hydrogels. Chem. Commun. 2017, 53 (23), 33503353,  DOI: 10.1039/C7CC00765E
  22. 22
    Hong, S. H.; Kim, S.; Park, J. P.; Shin, M.; Kim, K.; Ryu, J. H.; Lee, H. Dynamic Bonds between Boronic Acid and Alginate: Hydrogels with Stretchable, Self-Healing, Stimuli-Responsive, Remoldable, and Adhesive Properties. Biomacromolecules 2018, 19 (6), 20532061,  DOI: 10.1021/acs.biomac.8b00144
  23. 23
    Li, W. W.; Gao, F. X.; Wang, X. Q.; Zhang, N.; Ma, M. M. Strong and Robust Polyaniline-Based Supramolecular Hydrogels for Flexible Supercapacitors. Angew. Chem., Int. Ed. 2016, 55 (32), 91969201,  DOI: 10.1002/anie.201603417
  24. 24
    Casassa, E. Z.; Sarquis, A. M.; Vandyke, C. H. The Gelation of Polyvinyl-Alcohol with Borax - a Novel Class Participation Experiment Involving the Preparation and Properties of a Slime. J. Chem. Educ. 1986, 63 (1), 5760,  DOI: 10.1021/ed063p57
  25. 25
    Sinton, S. W. Complexation Chemistry of Sodium-Borate with Polyvinyl-Alcohol) and Small Diols - a B-11 NMR-Study. Macromolecules 1987, 20 (10), 24302441,  DOI: 10.1021/ma00176a018
  26. 26
    Wang, H. H.; Shyr, T. W.; Hu, M. S. The elastic property of polyvinyl alcohol gel with boric acid as a crosslinking agent. J. Am. Chem. Soc. 1999, 74 (13), 30463052,  DOI: 10.1002/(SICI)1097-4628(19991220)74:13<3046::AID-APP6>3.0.CO;2-1
  27. 27
    Duncan, T. T.; Berrie, B. H.; Weiss, R. G. Soft, Peelable Organogels from Partially Hydrolyzed Poly(vinyl acetate) and Benzene-1,4-diboronic Acid: Applications to Clean Works of Art. ACS Appl. Mater. Interfaces 2017, 9 (33), 2806928078,  DOI: 10.1021/acsami.7b09473
  28. 28
    Duncan, T. T.; Weiss, R. G. Influence of length and structure of aryl boronic acid crosslinkers on organogels with partially hydrolyzed poly(vinyl acetate). Colloid Polym. Sci. 2018, 296 (6), 10471056,  DOI: 10.1007/s00396-018-4326-7
  29. 29
    Nunes, M. A. P.; Gois, P. M. P.; Rosa, M. E.; Martins, S.; Fernandes, P. C. B.; Ribeiro, M. H. L. Boronic acids as efficient cross linkers for PVA: synthesis and application of tunable hollow microspheres in biocatalysis. Tetrahedron 2016, 72 (46), 72937305,  DOI: 10.1016/j.tet.2016.02.017
  30. 30
    Parris, M. D.; MacKay, B. A.; Rathke, J. W.; Klingler, R. J.; Gerald, R. E. Influence of Pressure on Boron Cross-Linked Polymer Gels. Macromolecules 2008, 41 (21), 81818186,  DOI: 10.1021/ma801187q
  31. 31
    Shim, J.; Lee, J. S.; Lee, J. H.; Kim, H. J.; Lee, J.-C. Gel Polymer Electrolytes Containing Anion-Trapping Boron Moieties for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2016, 8 (41), 2774027752,  DOI: 10.1021/acsami.6b09601
  32. 32
    Wang, X.; Liu, Z.; Kong, Q.; Jiang, W.; Yao, J.; Zhang, C.; Cui, G. A single-ion gel polymer electrolyte based on polymeric lithium tartaric acid borate and its superior battery performance. Solid State Ionics 2014, 262, 747753,  DOI: 10.1016/j.ssi.2013.09.007
  33. 33
    Dai, K.; Zheng, Y.; Wei, W. F. Organoboron-Containing Polymer Electrolytes for High-Performance Lithium Batteries. Adv. Funct. Mater. 2021, 31 (13), 2008632,  DOI: 10.1002/adfm.202008632
  34. 34
    Kato, Y.; Ishihara, T.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. A high electrode-reaction rate for high-power-density lithium-ion secondary batteries by the addition of a lewis acid. Angew. Chem., Int. Ed. 2004, 43 (15), 19661969,  DOI: 10.1002/anie.200353220
  35. 35
    Zhu, Y. S.; Wang, X. J.; Hou, Y. Y.; Gao, X. W.; Liu, L. L.; Wu, Y. P.; Shimizu, M. A new single-ion polymer electrolyte based on polyvinyl alcohol for lithium ion batteries. Electrochim. Acta 2013, 87, 113118,  DOI: 10.1016/j.electacta.2012.08.114
  36. 36
    Zhang, Y.; Cai, W.; Rohan, R.; Pan, M.; Liu, Y.; Liu, X.; Li, C.; Sun, Y.; Cheng, H. Toward ambient temperature operation with all-solid-state lithium metal batteries with a sp3 boron-based solid single ion conducting polymer electrolyte. J. Power Sources 2016, 306, 152161,  DOI: 10.1016/j.jpowsour.2015.12.010
  37. 37
    Sun, X.-G.; Liu, G.; Xie, J.; Han, Y.; Kerr, J. B. New gel polyelectrolytes for rechargeable lithium batteries. Solid State Ionics 2004, 175 (1), 713716,  DOI: 10.1016/j.ssi.2003.11.043
  38. 38
    Deng, K.; Wang, S.; Ren, S.; Han, D.; Xiao, M.; Meng, Y. Network type sp3 boron-based single-ion conducting polymer electrolytes for lithium ion batteries. J. Power Sources 2017, 360, 98105,  DOI: 10.1016/j.jpowsour.2017.06.006
  39. 39
    Guzmán-González, G.; Ávila-Paredes, H. J.; Rivera, E.; González, I. Electrochemical Characterization of Single Lithium-Ion Conducting Polymer Electrolytes Based on sp3 Boron and Poly(ethylene glycol) Bridges. ACS Appl. Mater. Interfaces 2018, 10 (36), 3024730256,  DOI: 10.1021/acsami.8b02519
  40. 40
    Van Humbeck, J. F.; Aubrey, M. L.; Alsbaiee, A.; Ameloot, R.; Coates, G. W.; Dichtel, W. R.; Long, J. R. Tetraarylborate polymer networks as single-ion conducting solid electrolytes. Chem. Sci. 2015, 6 (10), 54995505,  DOI: 10.1039/C5SC02052B
  41. 41
    Shim, J.; Kim, D.-G.; Kim, H. J.; Lee, J. H.; Lee, J.-C. Polymer Composite Electrolytes Having Core-Shell Silica Fillers with Anion-Trapping Boron Moiety in the Shell Layer for All-Solid-State Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7 (14), 76907701,  DOI: 10.1021/acsami.5b00618
  42. 42
    Li, S. B.; Zuo, C.; Zhang, Y.; Wang, J. R.; Gan, H. H.; Li, S. Q.; Yu, L. P.; Zhou, B. H.; Xue, Z. G. Covalently cross-linked polymer stabilized electrolytes with self-healing performance via boronic ester bonds. Polym. Chem. 2020, 11 (36), 58935902,  DOI: 10.1039/D0PY00728E
  43. 43
    Wang, L.; Yan, J.; Zhong, M.; Zhang, J.; Shen, W.; Guo, S. Regulating Lithium-Ion Transference Number of a Poly(vinyl alcohol)-Based Gel Electrolyte by the Incorporation of H3BO3 as an Anion Trapper. ACS Appl. Energy Mater. 2022, 5 (3), 28732880,  DOI: 10.1021/acsaem.1c03549
  44. 44
    Hong, D. G.; Baik, J.-H.; Kim, S.; Lee, J.-C. Solid polymer electrolytes based on polysiloxane with anion-trapping boron moieties for all-solid-state lithium metal batteries. Polymer 2022, 240, 124517,  DOI: 10.1016/j.polymer.2022.124517
  45. 45
    Guzman-Gonzalez, G.; Vauthier, S.; Alvarez-Tirado, M.; Cotte, S.; Castro, L.; Gueguen, A.; Casado, N.; Mecerreyes, D. Single-Ion Lithium Conducting Polymers with High Ionic Conductivity Based on Borate Pendant Groups. Angew. Chem., Int. Ed. 2022, 61 (8), e202114024,  DOI: 10.1002/anie.202114024
  46. 46
    Alvarez-Tirado, M.; Guzman-Gonzalez, G.; Vauthier, S.; Cotte, S.; Gueguen, A.; Castro, L.; Mecerreyes, D. Designing Boron-Based Single-Ion Gel Polymer Electrolytes for Lithium Batteries by Photopolymerization. Macromol. Chem. Phys. 2022, 223 (8), 2100407,  DOI: 10.1002/macp.202100407
  47. 47
    Chen, Q.; Shi, Y.; Sheng, K.; Zheng, J.; Xu, C. Dynamically Cross-Linked Hydrogel Electrolyte with Remarkable Stretchability and Self-Healing Capability for Flexible Electrochromic Devices. ACS Appl. Mater. Interfaces 2021, 13 (47), 5654456553,  DOI: 10.1021/acsami.1c15432
  48. 48
    Gregory, G. L.; Jenisch, L. M.; Charles, B.; Kociok-Köhn, G.; Buchard, A. Polymers from Sugars and CO2: Synthesis and Polymerization of a d-Mannose-Based Cyclic Carbonate. Macromolecules 2016, 49 (19), 71657169,  DOI: 10.1021/acs.macromol.6b01492
  49. 49
    Piccini, M.; Leak, D. J.; Chuck, C. J.; Buchard, A. Polymers from sugars and unsaturated fatty acids: ADMET polymerisation of monomers derived from d-xylose, d-mannose and castor oil. Polym. Chem. 2020, 11 (15), 26812691,  DOI: 10.1039/C9PY01809C
  50. 50
    McGuire, T. M.; Bowles, J.; Deane, E.; Farrar, E. H. E.; Grayson, M. N.; Buchard, A. Control of Crystallinity and Stereocomplexation of Synthetic Carbohydrate Polymers from d- and l-Xylose. Angew. Chem., Int. Ed. 2021, 60 (9), 45244528,  DOI: 10.1002/anie.202013562
  51. 51
    McGuire, T. M.; Clark, E. F.; Buchard, A. Polymers from Sugars and Cyclic Anhydrides: Ring-Opening Copolymerization of a d-Xylose Anhydrosugar Oxetane. Macromolecules 2021, 54 (11), 50945105,  DOI: 10.1021/acs.macromol.1c00365
  52. 52
    Gregory, G. L.; Kociok-Kohn, G.; Buchard, A. Polymers from sugars and CO2: ring-opening polymerisation and copolymerisation of cyclic carbonates derived from 2-deoxy-d-ribose. Polym. Chem. 2017, 8 (13), 20932104,  DOI: 10.1039/C7PY00236J
  53. 53
    Kumar, R.; Gao, W.; Gross, R. A. Functionalized Polylactides: Preparation and Characterization of [l]-Lactide-co-Pentofuranose. Macromolecules 2002, 35 (18), 68356844,  DOI: 10.1021/ma020041a
  54. 54
    Felder, S. E.; Redding, M. J.; Noel, A.; Grayson, S. M.; Wooley, K. L. Organocatalyzed ROP of a Glucopyranoside Derived Five-Membered Cyclic Carbonate. Macromolecules 2018, 51 (5), 17871797,  DOI: 10.1021/acs.macromol.7b01785
  55. 55
    Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A review. Prog. Polym. sci. 2010, 35 (5), 578622,  DOI: 10.1016/j.progpolymsci.2009.10.001
  56. 56
    Xiao, R.; Grinstaff, M. W. Chemical synthesis of polysaccharides and polysaccharide mimetics. Prog. Polym. sci. 2017, 74, 78116,  DOI: 10.1016/j.progpolymsci.2017.07.009
  57. 57
    Zhu, Y.; Romain, C.; Williams, C. K. Sustainable polymers from renewable resources. Nature 2016, 540 (7633), 354362,  DOI: 10.1038/nature21001
  58. 58
    Galbis, J. A.; García-Martín, M. d. G.; de Paz, M. V.; Galbis, E. Synthetic Polymers from Sugar-Based Monomers. Chem. Rev. 2016, 116 (3), 16001636,  DOI: 10.1021/acs.chemrev.5b00242
  59. 59
    Gregory, G. L.; Lopez-Vidal, E. M.; Buchard, A. Polymers from sugars: cyclic monomer synthesis, ring-opening polymerisation, material properties and applications. Chem. Commun. 2017, 53 (14), 21982217,  DOI: 10.1039/C6CC09578J
  60. 60
    Ma, L.; Jin, M.; Yan, C.; Guo, H.; Ma, X. Gel Polymer Electrolyte with Anion-Trapping Boron Moieties via One-Step Synthesis for Symmetrical Supercapacitors. Macromol. Mater. Eng. 2020, 305 (7), 1900807,  DOI: 10.1002/mame.201900807
  61. 61
    Dai, K.; Ma, C.; Feng, Y.; Zhou, L.; Kuang, G.; Zhang, Y.; Lai, Y.; Cui, X.; Wei, W. A borate-rich, cross-linked gel polymer electrolyte with near-single ion conduction for lithium metal batteries. J. Mater. Chem. A 2019, 7 (31), 1854718557,  DOI: 10.1039/C9TA05938E
  62. 62
    Zeng, X.; Dong, L.; Fu, J.; Chen, L.; Zhou, J.; Zong, P.; Liu, G.; Shi, L. Enhanced interfacial stability with a novel boron-centered crosslinked hybrid polymer gel electrolytes for lithium metal batteries. Chem. Eng. J. 2022, 428, 131100,  DOI: 10.1016/j.cej.2021.131100
  63. 63
    Chen, Q.; Shi, Y.; Sheng, K.; Zheng, J.; Xu, C. Dynamically Cross-Linked Hydrogel Electrolyte with Remarkable Stretchability and Self-Healing Capability for Flexible Electrochromic Devices. ACS Appl. Mater. Interfaces 2021, 13 (47), 5654456553,  DOI: 10.1021/acsami.1c15432
  64. 64
    Homann, G.; Stolz, L.; Winter, M.; Kasnatscheew, J. Elimination of ″Voltage Noise″ of Poly (Ethylene Oxide)-Based Solid Electrolytes in High-Voltage Lithium Batteries: Linear versus Network Polymers. Iscience 2020, 23 (6), 101225,  DOI: 10.1016/j.isci.2020.101225
  65. 65
    Gou, J. R.; Liu, W. Y.; Tang, A. M. A renewable gel polymer electrolyte based on the different sized carboxylated cellulose with satisfactory comprehensive performance for rechargeable lithium ion battery. Polymer 2020, 208, 122943,  DOI: 10.1016/j.polymer.2020.122943
  66. 66
    Gou, J. R.; Liu, W. Y.; Tang, A. M.; Xie, H. A phosphorylated nanocellulose/hydroxypropyl methylcellulose composite matrix: A biodegradable, flame-retardant and self-standing gel polymer electrolyte towards eco-friendly and high safety lithium ion batteries. Eur. Polym. J. 2021, 158, 110703,  DOI: 10.1016/j.eurpolymj.2021.110703
  67. 67
    Li, M. X.; Wang, X. W.; Yang, Y. Q.; Chang, Z.; Wu, Y. P.; Holze, R. A dense cellulose-based membrane as a renewable host for gel polymer electrolyte of lithium ion batteries. J. Membr. sci. 2015, 476, 112118,  DOI: 10.1016/j.memsci.2014.10.056
  68. 68
    Yang, H. Z.; Liu, Y.; Kong, L. B.; Kang, L.; Ran, F. Biopolymer-based carboxylated chitosan hydrogel film crosslinked by HCl as gel polymer electrolyte for all-solid-sate supercapacitors. J. Power Sources 2019, 426, 4754,  DOI: 10.1016/j.jpowsour.2019.04.023
  69. 69
    Kamisan, A. S.; Kudin, T. I. T.; Ali, A. M. M.; Yahya, M. Z. A. Electrical and physical studies on 49% methyl-grafted natural rubber-based composite polymer gel electrolytes. Electrochim. Acta 2011, 57, 207211,  DOI: 10.1016/j.electacta.2011.06.096
  70. 70
    Careem, M. A.; Noor, I. S. M.; Arof, A. K., Impedance Spectroscopy in Polymer Electrolyte Characterisation. In Polymer Electrolytes: Characterization Techniques and Energy Applications, 1st ed.; Winie, T., Arof, A. K., Thomas, S., Eds.; Wiley-VCH verlag GmbH & Co.: Weinheim Germany, 2020; pp 2364.

Cited By

Click to copy section linkSection link copied!

This article is cited by 15 publications.

  1. Jean G. A. Ruthes, Stefanie Arnold, Kaitlyn Prenger, Ana C. Jaski, Vanessa Klobukoski, Izabel C. Riegel-Vidotti, Volker Presser. Functional Gel-Based Electrochemical Energy Storage. Chemistry of Materials 2024, 36 (12) , 5861-5877. https://doi.org/10.1021/acs.chemmater.4c00321
  2. Ekata Saha, Andrea Ruiu, Savannah C. Zacharias, Anna Ström, Henrik Sundén. Dual Role of Boronic Acid Modification in Enhancing Rheological and Photophysical Properties of Supramolecular Gels. Macromolecules 2024, 57 (7) , 3183-3189. https://doi.org/10.1021/acs.macromol.3c02091
  3. Matthew Oshinowo, Marco Piccini, Gabriele Kociok-Köhn, Frank Marken, Antoine Buchard. Xylose- and Nucleoside-Based Polymers via Thiol–ene Polymerization toward Sugar-Derived Solid Polymer Electrolytes. ACS Applied Polymer Materials 2024, 6 (3) , 1622-1632. https://doi.org/10.1021/acsapm.3c02119
  4. Indrajit Pal, Santanu Majumdar, Gerald Lepcha, Kazi Tawsif Ahmed, Suresh Kumar Yatirajula, Subhratanu Bhattacharya, Rudra Chakravarti, Bireswar Bhattacharya, Swadesh Ranjan Biswas, Biswajit Dey. Exploration of Variable Solvent Directed Self-Healable Supramolecular M(II)-Metallogels (M = Co, Ni, Zn) of Azelaic Acid: Investigating Temperature-Dependent Ion Conductivity and Antibacterial Efficiency. ACS Applied Bio Materials 2023, 6 (12) , 5442-5457. https://doi.org/10.1021/acsabm.3c00675
  5. Amir Khan, Ravinder Reddy Kisannagar, Sadiq Mahmood, Wei-Tsung Chuang, Monica Katiyar, Dipti Gupta, Hong-Cheu Lin. Intrinsically Stretchable Conductive Self-Healable Organogels for Strain, Pressure, Temperature, and Humidity Sensing. ACS Applied Materials & Interfaces 2023, 15 (36) , 42954-42964. https://doi.org/10.1021/acsami.3c08111
  6. Yidong Xu, Yao Sun, Zewei Yao, Yinmao Wei. Hierarchical Mesoporous Metal–Organic Frameworks with Boric Acid Sites on the Inner Surface of Small Mesopores for the Extraction of Nucleotides in Human Plasma Samples. ACS Applied Materials & Interfaces 2023, 15 (25) , 30643-30652. https://doi.org/10.1021/acsami.3c05025
  7. Daun Jeong, Da‐Sol Kwon, Gwangbin Won, Seunghyeon Kim, Joona Bang, Jimin Shim. Toward Sustainable Polymer Materials for Rechargeable Batteries: Utilizing Natural Feedstocks and Recycling/Upcycling of Polymer Waste. ChemSusChem 2024, 17 (22) https://doi.org/10.1002/cssc.202401010
  8. Ella F. Clark, Alexandra Howard, Sebastian D. Morales Feliu, James F. McCabe, Jonathan C. Burley, Vincenzo Taresco, Antoine Buchard. d -Xylose oxetane copolymers as bioderived and tuneable polyesters for amorphous solid dispersions. RSC Applied Polymers 2024, 2 (6) , 1104-1112. https://doi.org/10.1039/D4LP00203B
  9. Ashok Kumar Jangid, Kyobum Kim. Phenylboronic acid-functionalized biomaterials for improved cancer immunotherapy via sialic acid targeting. Advances in Colloid and Interface Science 2024, 333 , 103301. https://doi.org/10.1016/j.cis.2024.103301
  10. James R. Runge, Bethan Davies, Antoine Buchard. Xylose-derived thionocarbamates as a synthetic handle towards a functional platform of sugar-based polymers. Polymer Chemistry 2024, 15 (31) , 3149-3156. https://doi.org/10.1039/D4PY00540F
  11. Sofia Nieves Casillas‐Popova, Nishadi Dilkushi Lokuge, Brandon Andrade‐Gagnon, Farhan Rahman Chowdhury, Cameron D. Skinner, Brandon L. Findlay, Jung Kwon Oh. pH‐Responsive Degradable Electro‐Spun Nanofibers Crosslinked via Boronic Ester Chemistry for Smart Wound Dressings. Macromolecular Bioscience 2024, 50 https://doi.org/10.1002/mabi.202400217
  12. Kieran G. Stakem, Freddie J. Leslie, Georgina L. Gregory. Polymer design for solid-state batteries and wearable electronics. Chemical Science 2024, 15 (27) , 10281-10307. https://doi.org/10.1039/D4SC02501F
  13. Yingshan Ma, Sofia M. Morozova, Eugenia Kumacheva. From Nature‐Sourced Polysaccharide Particles to Advanced Functional Materials. Advanced Materials 2024, 36 (23) https://doi.org/10.1002/adma.202312707
  14. Mahak Fatima, Waleed H. Almalki, Tasneem Khan, Amirhossein Sahebkar, Prashant Kesharwani. Harnessing the Power of Stimuli‐Responsive Nanoparticles as an Effective Therapeutic Drug Delivery System. Advanced Materials 2024, 36 (24) https://doi.org/10.1002/adma.202312939
  15. Jiao Guo, Hongliang Xu, Yuxue Sun, Kai Chen, Xiaorong Zhang, Haiming Xie, Ying Jiang, Jun Liu. Borate-containing triblock copolymer electrolytes for improved lithium-ion transference number and interface stability. Journal of Colloid and Interface Science 2024, 660 , 565-573. https://doi.org/10.1016/j.jcis.2024.01.097

ACS Applied Energy Materials

Cite this: ACS Appl. Energy Mater. 2023, 6, 5, 2924–2935
Click to copy citationCitation copied!
https://doi.org/10.1021/acsaem.2c03937
Published February 21, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

4444

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Synthesis of Monomer d-1, Subsequent Anionic Ring-Opening Polymerization, and Post-Polymerization Acetal Deprotection to Yield Hydroxy-Polymer, dp-poly(d-1)

    Scheme 2

    Scheme 2. Cross-Linking of dp-poly(d-1) and 1,4-Phenylenediboronic Acid (PDBA)

    Figure 1

    Figure 1. Vial inversion tests of (a) a 24 mg mL–1 solution of PDBA in DMSO, (b) a 0.287 mol L–1 solution of dp-poly(d-1), and (c) a reaction between dp-poly(d-1) and PDBA observed over several hours. (d) Phase diagram of the dp-poly(d-1)/PDBA system in DMSO at room temperature.

    Figure 2

    Figure 2. FT-IR spectra of dp-poly(d-1) (black), PDBA (red), and organogel (blue) made from dp-poly(d-1) (0.287 mol L–1) and 0.5 equiv of PDBA (with respect to monomer unit).

    Figure 3

    Figure 3. (a) Dynamic frequency sweep of organogels made from dp-poly(d-1) (0.287 mol L–1) and increasing PDBA equiv. (b) Dynamic frequency sweep of organogels made from dp-poly(d-1) (increasing concentrations) and PDBA (1 equiv).

    Figure 4

    Figure 4. Change in moduli during a strain ramp of an organogel (0.144 mol L–1 dp-poly(d-1), 0.50 equiv of PDBA) measured at a frequency of 1 Hz. The percentage strain was increased from 0.007% to 100%; then strain was released; and the moduli were measured at 0.007% strain.

    Figure 5

    Figure 5. Gel fracture and self-healing in sequential strain steps. Top shows the measured storage moduli at each strain step, as described in the bottom plot (0% → 10% → 0% → 20% → 0% → 40% → 0% → 60% → 0% → 80% → 0% → 100% → 0%, measured at 1 Hz).

    Figure 6

    Figure 6. FE-SEM imaging of lyophilized organogels made from 0.144 M solutions of dp-poly(d-1) in DMSO with 0.25–1.00 equiv of PDBA.

    Figure 7

    Figure 7. TGA of (a) lyophilized and (b) nonlyophilized organogel samples made from dp-poly(d-1) and PDBA with a polymer concentration of 0.144 mol L–1 and PDBA equiv of 0.25–1.00. Annotations of the Td,max values of each sample.

    Scheme 3

    Scheme 3. Cross-Linking of dp-poly(d-1) and PDBA in the Presence of a Lithium Salt (LiTFSI)

    Figure 8

    Figure 8. Temperature dependence of ionic conductivity (σ) for lithiated (1 equiv of LiTFSI, 0.144 mol L–1; Li-gel) and nonlithiated (control) organogels made from dp-poly(d-1) (0.287 mol L–1 with respect to the repeating unit) with 0.50 equiv of PDBA.

    Figure 9

    Figure 9. Li+ stripping/plating voltage profile of Li-gel in a symmetrical Li|Li-gel|Li cell at 25 °C. The half-cycle duration was 1 h with a cutoff voltage of ±1.0 V.

  • References


    This article references 70 other publications.

    1. 1
      Asher, S. A.; Alexeev, V. L.; Goponenko, A. V.; Sharma, A. C.; Lednev, I. K.; Wilcox, C. S.; Finegold, D. N. Photonic Crystal Carbohydrate Sensors: Low Ionic Strength Sugar Sensing. J. Am. Chem. Soc. 2003, 125 (11), 33223329,  DOI: 10.1021/ja021037h
    2. 2
      Lerner, M. B.; Kybert, N.; Mendoza, R.; Villechenon, R.; Bonilla Lopez, M. A.; Charlie Johnson, A. T. Scalable, non-invasive glucose sensor based on boronic acid functionalized carbon nanotube transistors. Appl. Phys. Lett. 2013, 102 (18), 183113,  DOI: 10.1063/1.4804438
    3. 3
      Nishiyabu, R.; Kobayashi, H.; Kubo, Y. Dansyl-containing boronate hydrogel film as fluorescent chemosensor of copper ions in water. RSC Adv. 2012, 2 (16), 65556561,  DOI: 10.1039/c2ra20516e
    4. 4
      Awino, J. K.; Gunasekara, R. W.; Zhao, Y. Selective Recognition of d-Aldohexoses in Water by Boronic Acid-Functionalized, Molecularly Imprinted Cross-Linked Micelles. J. Am. Chem. Soc. 2016, 138 (31), 97599762,  DOI: 10.1021/jacs.6b04613
    5. 5
      Chen, W.; Guo, Z.; Yu, H.; Liu, Q.; Fu, M. Molecularly imprinted colloidal array with multi-boronic acid sites for glycoprotein detection under neutral pH. J. Colloid Interface Sci. 2022, 607, 11631172,  DOI: 10.1016/j.jcis.2021.09.048
    6. 6
      He, P.; Zhu, H.; Ma, Y.; Liu, N.; Niu, X.; Wei, M.; Pan, J. Rational design and fabrication of surface molecularly imprinted polymers based on multi-boronic acid sites for selective capture glycoproteins. Chem. Eng. J. 2019, 367, 5563,  DOI: 10.1016/j.cej.2019.02.140
    7. 7
      Zhao, Z. W.; Yao, X. M.; Zhang, Z.; Chen, L.; He, C. L.; Chen, X. S. Boronic Acid Shell-Crosslinked Dextran-b-PLA Micelles for Acid-Responsive Drug Delivery. Macromol. Biosci. 2014, 14 (11), 16091618,  DOI: 10.1002/mabi.201400251
    8. 8
      Li, Y. P.; Xiao, W. W.; Xiao, K.; Berti, L.; Luo, J. T.; Tseng, H. P.; Fung, G.; Lam, K. S. Well-Defined, Reversible Boronate Crosslinked Nanocarriers for Targeted Drug Delivery in Response to Acidic pH Values and cis-Diols. Angew. Chem., Int. Ed. 2012, 51 (12), 28642869,  DOI: 10.1002/anie.201107144
    9. 9
      Su, J.; Chen, F.; Cryns, V. L.; Messersmith, P. B. Catechol Polymers for pH-Responsive, Targeted Drug Delivery to Cancer Cells. J. Am. Chem. Soc. 2011, 133 (31), 1185011853,  DOI: 10.1021/ja203077x
    10. 10
      Jiang, H.-P.; Qi, C.-B.; Chu, J.-M.; Yuan, B.-F.; Feng, Y.-Q. Profiling of cis-Diol-containing Nucleosides and Ribosylated Metabolites by Boronate-affinity Organic-silica Hybrid Monolithic Capillary Liquid Chromatography/Mass Spectrometry. Sci. Rep. 2015, 5 (1), 7785,  DOI: 10.1038/srep07785
    11. 11
      Koyama, T.; Terauchi, K.-i. Synthesis and application of boronic acid-immobilized porous polymer particles: a novel packing for high-performance liquid affinity chromatography. J. Chromatogr. B: Biomed. Sci. Appl. 1996, 679 (1), 3140,  DOI: 10.1016/0378-4347(96)00006-0
    12. 12
      Brooks, W. L. A.; Sumerlin, B. S. Synthesis and Applications of Boronic Acid-Containing Polymers: From Materials to Medicine. Chem. Rev. 2016, 116 (3), 13751397,  DOI: 10.1021/acs.chemrev.5b00300
    13. 13
      He, L. H.; Fullenkamp, D. E.; Rivera, J. G.; Messersmith, P. B. pH responsive self-healing hydrogels formed by boronate-catechol complexation. Chem. Commun. 2011, 47 (26), 74977499,  DOI: 10.1039/c1cc11928a
    14. 14
      Figueiredo, T.; Cosenza, V.; Ogawa, Y.; Jeacomine, I.; Vallet, A.; Ortega, S.; Michel, R.; Olsson, J. D. M.; Gerfaud, T.; Boiteau, J. G.; Jing, J.; Harris, C.; Auzely-Velty, R. Boronic acid and diol-containing polymers: how to choose the correct couple to form ″strong″ hydrogels at physiological pH. Soft Matter 2020, 16 (15), 36283641,  DOI: 10.1039/D0SM00178C
    15. 15
      Brooks, W. L. A.; Deng, C. C.; Sumerlin, B. S. Structure-Reactivity Relationships in Boronic Acid-Diol Complexation. ACS Omega 2018, 3 (12), 1786317870,  DOI: 10.1021/acsomega.8b02999
    16. 16
      Deng, C. C.; Brooks, W. L. A.; Abboud, K. A.; Sumerlin, B. S. Boronic Acid-Based Hydrogels Undergo Self-Healing at Neutral and Acidic pH. ACS Macro Lett. 2015, 4 (2), 220224,  DOI: 10.1021/acsmacrolett.5b00018
    17. 17
      Smithmyer, M. E.; Deng, C. C.; Cassel, S. E.; LeValley, P. J.; Sumerlin, B. S.; Kloxin, A. M. Self-Healing Boronic Acid-Based Hydrogels for 3D Co-cultures. ACS Macro Lett. 2018, 7 (9), 11051110,  DOI: 10.1021/acsmacrolett.8b00462
    18. 18
      Wang, X.-T.; Deng, X.; Zhang, T.-D.; Zhang, X.; Shi, W.-P.; Lai, J.; Zhou, H.; Ye, Y.-J.; Zhang, C.-Y.; Yin, D.-C. Biocompatible self-healing hydrogels based on boronic acid-functionalized polymer and laponite nanocomposite for water pollutant removal. Environ. Chem. Lett. 2022, 20 (1), 8190,  DOI: 10.1007/s10311-021-01350-4
    19. 19
      Huang, Z. J.; Delparastan, P.; Burch, P.; Cheng, J.; Cao, Y.; Messersmith, P. B. Injectable dynamic covalent hydrogels of boronic acid polymers cross-linked by bioactive plant-derived polyphenols. Biomater. Sci. 2018, 6 (9), 24872495,  DOI: 10.1039/C8BM00453F
    20. 20
      Dong, Y. Z.; Wang, W. H.; Veiseh, O.; Appel, E. A.; Xue, K.; Webber, M. J.; Tang, B. C.; Yang, X. W.; Weir, G. C.; Langer, R.; Anderson, D. G. Injectable and Glucose-Responsive Hydrogels Based on Boronic Acid-Glucose Complexation. Langmuir 2016, 32 (34), 87438747,  DOI: 10.1021/acs.langmuir.5b04755
    21. 21
      Pettignano, A.; Grijalvo, S.; Haring, M.; Eritja, R.; Tanchoux, N.; Quignard, F.; Diaz, D. D. Boronic acid-modified alginate enables direct formation of injectable, self-healing and multistimuli-responsive hydrogels. Chem. Commun. 2017, 53 (23), 33503353,  DOI: 10.1039/C7CC00765E
    22. 22
      Hong, S. H.; Kim, S.; Park, J. P.; Shin, M.; Kim, K.; Ryu, J. H.; Lee, H. Dynamic Bonds between Boronic Acid and Alginate: Hydrogels with Stretchable, Self-Healing, Stimuli-Responsive, Remoldable, and Adhesive Properties. Biomacromolecules 2018, 19 (6), 20532061,  DOI: 10.1021/acs.biomac.8b00144
    23. 23
      Li, W. W.; Gao, F. X.; Wang, X. Q.; Zhang, N.; Ma, M. M. Strong and Robust Polyaniline-Based Supramolecular Hydrogels for Flexible Supercapacitors. Angew. Chem., Int. Ed. 2016, 55 (32), 91969201,  DOI: 10.1002/anie.201603417
    24. 24
      Casassa, E. Z.; Sarquis, A. M.; Vandyke, C. H. The Gelation of Polyvinyl-Alcohol with Borax - a Novel Class Participation Experiment Involving the Preparation and Properties of a Slime. J. Chem. Educ. 1986, 63 (1), 5760,  DOI: 10.1021/ed063p57
    25. 25
      Sinton, S. W. Complexation Chemistry of Sodium-Borate with Polyvinyl-Alcohol) and Small Diols - a B-11 NMR-Study. Macromolecules 1987, 20 (10), 24302441,  DOI: 10.1021/ma00176a018
    26. 26
      Wang, H. H.; Shyr, T. W.; Hu, M. S. The elastic property of polyvinyl alcohol gel with boric acid as a crosslinking agent. J. Am. Chem. Soc. 1999, 74 (13), 30463052,  DOI: 10.1002/(SICI)1097-4628(19991220)74:13<3046::AID-APP6>3.0.CO;2-1
    27. 27
      Duncan, T. T.; Berrie, B. H.; Weiss, R. G. Soft, Peelable Organogels from Partially Hydrolyzed Poly(vinyl acetate) and Benzene-1,4-diboronic Acid: Applications to Clean Works of Art. ACS Appl. Mater. Interfaces 2017, 9 (33), 2806928078,  DOI: 10.1021/acsami.7b09473
    28. 28
      Duncan, T. T.; Weiss, R. G. Influence of length and structure of aryl boronic acid crosslinkers on organogels with partially hydrolyzed poly(vinyl acetate). Colloid Polym. Sci. 2018, 296 (6), 10471056,  DOI: 10.1007/s00396-018-4326-7
    29. 29
      Nunes, M. A. P.; Gois, P. M. P.; Rosa, M. E.; Martins, S.; Fernandes, P. C. B.; Ribeiro, M. H. L. Boronic acids as efficient cross linkers for PVA: synthesis and application of tunable hollow microspheres in biocatalysis. Tetrahedron 2016, 72 (46), 72937305,  DOI: 10.1016/j.tet.2016.02.017
    30. 30
      Parris, M. D.; MacKay, B. A.; Rathke, J. W.; Klingler, R. J.; Gerald, R. E. Influence of Pressure on Boron Cross-Linked Polymer Gels. Macromolecules 2008, 41 (21), 81818186,  DOI: 10.1021/ma801187q
    31. 31
      Shim, J.; Lee, J. S.; Lee, J. H.; Kim, H. J.; Lee, J.-C. Gel Polymer Electrolytes Containing Anion-Trapping Boron Moieties for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2016, 8 (41), 2774027752,  DOI: 10.1021/acsami.6b09601
    32. 32
      Wang, X.; Liu, Z.; Kong, Q.; Jiang, W.; Yao, J.; Zhang, C.; Cui, G. A single-ion gel polymer electrolyte based on polymeric lithium tartaric acid borate and its superior battery performance. Solid State Ionics 2014, 262, 747753,  DOI: 10.1016/j.ssi.2013.09.007
    33. 33
      Dai, K.; Zheng, Y.; Wei, W. F. Organoboron-Containing Polymer Electrolytes for High-Performance Lithium Batteries. Adv. Funct. Mater. 2021, 31 (13), 2008632,  DOI: 10.1002/adfm.202008632
    34. 34
      Kato, Y.; Ishihara, T.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. A high electrode-reaction rate for high-power-density lithium-ion secondary batteries by the addition of a lewis acid. Angew. Chem., Int. Ed. 2004, 43 (15), 19661969,  DOI: 10.1002/anie.200353220
    35. 35
      Zhu, Y. S.; Wang, X. J.; Hou, Y. Y.; Gao, X. W.; Liu, L. L.; Wu, Y. P.; Shimizu, M. A new single-ion polymer electrolyte based on polyvinyl alcohol for lithium ion batteries. Electrochim. Acta 2013, 87, 113118,  DOI: 10.1016/j.electacta.2012.08.114
    36. 36
      Zhang, Y.; Cai, W.; Rohan, R.; Pan, M.; Liu, Y.; Liu, X.; Li, C.; Sun, Y.; Cheng, H. Toward ambient temperature operation with all-solid-state lithium metal batteries with a sp3 boron-based solid single ion conducting polymer electrolyte. J. Power Sources 2016, 306, 152161,  DOI: 10.1016/j.jpowsour.2015.12.010
    37. 37
      Sun, X.-G.; Liu, G.; Xie, J.; Han, Y.; Kerr, J. B. New gel polyelectrolytes for rechargeable lithium batteries. Solid State Ionics 2004, 175 (1), 713716,  DOI: 10.1016/j.ssi.2003.11.043
    38. 38
      Deng, K.; Wang, S.; Ren, S.; Han, D.; Xiao, M.; Meng, Y. Network type sp3 boron-based single-ion conducting polymer electrolytes for lithium ion batteries. J. Power Sources 2017, 360, 98105,  DOI: 10.1016/j.jpowsour.2017.06.006
    39. 39
      Guzmán-González, G.; Ávila-Paredes, H. J.; Rivera, E.; González, I. Electrochemical Characterization of Single Lithium-Ion Conducting Polymer Electrolytes Based on sp3 Boron and Poly(ethylene glycol) Bridges. ACS Appl. Mater. Interfaces 2018, 10 (36), 3024730256,  DOI: 10.1021/acsami.8b02519
    40. 40
      Van Humbeck, J. F.; Aubrey, M. L.; Alsbaiee, A.; Ameloot, R.; Coates, G. W.; Dichtel, W. R.; Long, J. R. Tetraarylborate polymer networks as single-ion conducting solid electrolytes. Chem. Sci. 2015, 6 (10), 54995505,  DOI: 10.1039/C5SC02052B
    41. 41
      Shim, J.; Kim, D.-G.; Kim, H. J.; Lee, J. H.; Lee, J.-C. Polymer Composite Electrolytes Having Core-Shell Silica Fillers with Anion-Trapping Boron Moiety in the Shell Layer for All-Solid-State Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7 (14), 76907701,  DOI: 10.1021/acsami.5b00618
    42. 42
      Li, S. B.; Zuo, C.; Zhang, Y.; Wang, J. R.; Gan, H. H.; Li, S. Q.; Yu, L. P.; Zhou, B. H.; Xue, Z. G. Covalently cross-linked polymer stabilized electrolytes with self-healing performance via boronic ester bonds. Polym. Chem. 2020, 11 (36), 58935902,  DOI: 10.1039/D0PY00728E
    43. 43
      Wang, L.; Yan, J.; Zhong, M.; Zhang, J.; Shen, W.; Guo, S. Regulating Lithium-Ion Transference Number of a Poly(vinyl alcohol)-Based Gel Electrolyte by the Incorporation of H3BO3 as an Anion Trapper. ACS Appl. Energy Mater. 2022, 5 (3), 28732880,  DOI: 10.1021/acsaem.1c03549
    44. 44
      Hong, D. G.; Baik, J.-H.; Kim, S.; Lee, J.-C. Solid polymer electrolytes based on polysiloxane with anion-trapping boron moieties for all-solid-state lithium metal batteries. Polymer 2022, 240, 124517,  DOI: 10.1016/j.polymer.2022.124517
    45. 45
      Guzman-Gonzalez, G.; Vauthier, S.; Alvarez-Tirado, M.; Cotte, S.; Castro, L.; Gueguen, A.; Casado, N.; Mecerreyes, D. Single-Ion Lithium Conducting Polymers with High Ionic Conductivity Based on Borate Pendant Groups. Angew. Chem., Int. Ed. 2022, 61 (8), e202114024,  DOI: 10.1002/anie.202114024
    46. 46
      Alvarez-Tirado, M.; Guzman-Gonzalez, G.; Vauthier, S.; Cotte, S.; Gueguen, A.; Castro, L.; Mecerreyes, D. Designing Boron-Based Single-Ion Gel Polymer Electrolytes for Lithium Batteries by Photopolymerization. Macromol. Chem. Phys. 2022, 223 (8), 2100407,  DOI: 10.1002/macp.202100407
    47. 47
      Chen, Q.; Shi, Y.; Sheng, K.; Zheng, J.; Xu, C. Dynamically Cross-Linked Hydrogel Electrolyte with Remarkable Stretchability and Self-Healing Capability for Flexible Electrochromic Devices. ACS Appl. Mater. Interfaces 2021, 13 (47), 5654456553,  DOI: 10.1021/acsami.1c15432
    48. 48
      Gregory, G. L.; Jenisch, L. M.; Charles, B.; Kociok-Köhn, G.; Buchard, A. Polymers from Sugars and CO2: Synthesis and Polymerization of a d-Mannose-Based Cyclic Carbonate. Macromolecules 2016, 49 (19), 71657169,  DOI: 10.1021/acs.macromol.6b01492
    49. 49
      Piccini, M.; Leak, D. J.; Chuck, C. J.; Buchard, A. Polymers from sugars and unsaturated fatty acids: ADMET polymerisation of monomers derived from d-xylose, d-mannose and castor oil. Polym. Chem. 2020, 11 (15), 26812691,  DOI: 10.1039/C9PY01809C
    50. 50
      McGuire, T. M.; Bowles, J.; Deane, E.; Farrar, E. H. E.; Grayson, M. N.; Buchard, A. Control of Crystallinity and Stereocomplexation of Synthetic Carbohydrate Polymers from d- and l-Xylose. Angew. Chem., Int. Ed. 2021, 60 (9), 45244528,  DOI: 10.1002/anie.202013562
    51. 51
      McGuire, T. M.; Clark, E. F.; Buchard, A. Polymers from Sugars and Cyclic Anhydrides: Ring-Opening Copolymerization of a d-Xylose Anhydrosugar Oxetane. Macromolecules 2021, 54 (11), 50945105,  DOI: 10.1021/acs.macromol.1c00365
    52. 52
      Gregory, G. L.; Kociok-Kohn, G.; Buchard, A. Polymers from sugars and CO2: ring-opening polymerisation and copolymerisation of cyclic carbonates derived from 2-deoxy-d-ribose. Polym. Chem. 2017, 8 (13), 20932104,  DOI: 10.1039/C7PY00236J
    53. 53
      Kumar, R.; Gao, W.; Gross, R. A. Functionalized Polylactides: Preparation and Characterization of [l]-Lactide-co-Pentofuranose. Macromolecules 2002, 35 (18), 68356844,  DOI: 10.1021/ma020041a
    54. 54
      Felder, S. E.; Redding, M. J.; Noel, A.; Grayson, S. M.; Wooley, K. L. Organocatalyzed ROP of a Glucopyranoside Derived Five-Membered Cyclic Carbonate. Macromolecules 2018, 51 (5), 17871797,  DOI: 10.1021/acs.macromol.7b01785
    55. 55
      Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A review. Prog. Polym. sci. 2010, 35 (5), 578622,  DOI: 10.1016/j.progpolymsci.2009.10.001
    56. 56
      Xiao, R.; Grinstaff, M. W. Chemical synthesis of polysaccharides and polysaccharide mimetics. Prog. Polym. sci. 2017, 74, 78116,  DOI: 10.1016/j.progpolymsci.2017.07.009
    57. 57
      Zhu, Y.; Romain, C.; Williams, C. K. Sustainable polymers from renewable resources. Nature 2016, 540 (7633), 354362,  DOI: 10.1038/nature21001
    58. 58
      Galbis, J. A.; García-Martín, M. d. G.; de Paz, M. V.; Galbis, E. Synthetic Polymers from Sugar-Based Monomers. Chem. Rev. 2016, 116 (3), 16001636,  DOI: 10.1021/acs.chemrev.5b00242
    59. 59
      Gregory, G. L.; Lopez-Vidal, E. M.; Buchard, A. Polymers from sugars: cyclic monomer synthesis, ring-opening polymerisation, material properties and applications. Chem. Commun. 2017, 53 (14), 21982217,  DOI: 10.1039/C6CC09578J
    60. 60
      Ma, L.; Jin, M.; Yan, C.; Guo, H.; Ma, X. Gel Polymer Electrolyte with Anion-Trapping Boron Moieties via One-Step Synthesis for Symmetrical Supercapacitors. Macromol. Mater. Eng. 2020, 305 (7), 1900807,  DOI: 10.1002/mame.201900807
    61. 61
      Dai, K.; Ma, C.; Feng, Y.; Zhou, L.; Kuang, G.; Zhang, Y.; Lai, Y.; Cui, X.; Wei, W. A borate-rich, cross-linked gel polymer electrolyte with near-single ion conduction for lithium metal batteries. J. Mater. Chem. A 2019, 7 (31), 1854718557,  DOI: 10.1039/C9TA05938E
    62. 62
      Zeng, X.; Dong, L.; Fu, J.; Chen, L.; Zhou, J.; Zong, P.; Liu, G.; Shi, L. Enhanced interfacial stability with a novel boron-centered crosslinked hybrid polymer gel electrolytes for lithium metal batteries. Chem. Eng. J. 2022, 428, 131100,  DOI: 10.1016/j.cej.2021.131100
    63. 63
      Chen, Q.; Shi, Y.; Sheng, K.; Zheng, J.; Xu, C. Dynamically Cross-Linked Hydrogel Electrolyte with Remarkable Stretchability and Self-Healing Capability for Flexible Electrochromic Devices. ACS Appl. Mater. Interfaces 2021, 13 (47), 5654456553,  DOI: 10.1021/acsami.1c15432
    64. 64
      Homann, G.; Stolz, L.; Winter, M.; Kasnatscheew, J. Elimination of ″Voltage Noise″ of Poly (Ethylene Oxide)-Based Solid Electrolytes in High-Voltage Lithium Batteries: Linear versus Network Polymers. Iscience 2020, 23 (6), 101225,  DOI: 10.1016/j.isci.2020.101225
    65. 65
      Gou, J. R.; Liu, W. Y.; Tang, A. M. A renewable gel polymer electrolyte based on the different sized carboxylated cellulose with satisfactory comprehensive performance for rechargeable lithium ion battery. Polymer 2020, 208, 122943,  DOI: 10.1016/j.polymer.2020.122943
    66. 66
      Gou, J. R.; Liu, W. Y.; Tang, A. M.; Xie, H. A phosphorylated nanocellulose/hydroxypropyl methylcellulose composite matrix: A biodegradable, flame-retardant and self-standing gel polymer electrolyte towards eco-friendly and high safety lithium ion batteries. Eur. Polym. J. 2021, 158, 110703,  DOI: 10.1016/j.eurpolymj.2021.110703
    67. 67
      Li, M. X.; Wang, X. W.; Yang, Y. Q.; Chang, Z.; Wu, Y. P.; Holze, R. A dense cellulose-based membrane as a renewable host for gel polymer electrolyte of lithium ion batteries. J. Membr. sci. 2015, 476, 112118,  DOI: 10.1016/j.memsci.2014.10.056
    68. 68
      Yang, H. Z.; Liu, Y.; Kong, L. B.; Kang, L.; Ran, F. Biopolymer-based carboxylated chitosan hydrogel film crosslinked by HCl as gel polymer electrolyte for all-solid-sate supercapacitors. J. Power Sources 2019, 426, 4754,  DOI: 10.1016/j.jpowsour.2019.04.023
    69. 69
      Kamisan, A. S.; Kudin, T. I. T.; Ali, A. M. M.; Yahya, M. Z. A. Electrical and physical studies on 49% methyl-grafted natural rubber-based composite polymer gel electrolytes. Electrochim. Acta 2011, 57, 207211,  DOI: 10.1016/j.electacta.2011.06.096
    70. 70
      Careem, M. A.; Noor, I. S. M.; Arof, A. K., Impedance Spectroscopy in Polymer Electrolyte Characterisation. In Polymer Electrolytes: Characterization Techniques and Energy Applications, 1st ed.; Winie, T., Arof, A. K., Thomas, S., Eds.; Wiley-VCH verlag GmbH & Co.: Weinheim Germany, 2020; pp 2364.
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsaem.2c03937.

    • Cross-linking procedures, NMR spectra, photographs of vial inversion tests, FT-IR spectra, detailed rheological characterizations, FE-SEM microscopy, TGA and DSC traces, and additional electrochemistry results (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.