Next Article in Journal
Using Heating and Cooling Presses in Combination to Optimize the Consolidation Process of Polycarbonate-Based Unidirectional Thermoplastic Composite Tapes
Next Article in Special Issue
Advances in the Production of Sustainable Bacterial Nanocellulose from Banana Leaves
Previous Article in Journal
Mussel-Inspired Calcium Alginate/Polyacrylamide Dual Network Hydrogel: A Physical Barrier to Prevent Postoperative Re-Adhesion
Previous Article in Special Issue
Depolymerisation of Kraft Lignin by Tailor-Made Alkaliphilic Fungal Laccases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, and Proton Conductivity of Muconic Acid-Based Polyamides Bearing Sulfonated Moieties

by
Carlos Corona-García
1,
Alejandro Onchi
1,
Arlette A. Santiago
2,
Tania E. Soto
3,
Salomón Ramiro Vásquez-García
4,
Daniella Esperanza Pacheco-Catalán
5 and
Joel Vargas
1,*
1
Instituto de Investigaciones en Materiales, Unidad Morelia, Universidad Nacional Autónoma de México, Antigua Carretera a Pátzcuaro No. 8701, Col. Ex Hacienda de San José de la Huerta, Morelia C.P. 58190, Michoacán, Mexico
2
Escuela Nacional de Estudios Superiores, Unidad Morelia, Universidad Nacional Autónoma de México, Antigua Carretera a Pátzcuaro No. 8701, Col. Ex Hacienda de San José de la Huerta, Morelia C.P. 58190, Michoacán, Mexico
3
Centro de Investigaciones Químicas, Instituto de Investigación en Ciencias Básicas y Aplicadas, Universidad Autónoma del Estado de Morelos, Av. Universidad 1001, Cuernavaca C.P. 62209, Morelos, Mexico
4
Facultad de Ingeniería Química, Universidad Michoacana de San Nicolás de Hidalgo, General Francisco J. Múgica s/n, Morelia C.P. 58060, Michoacán, Mexico
5
Unidad de Energía Renovable, Centro de Investigación Científica de Yucatán, A.C. Carretera Sierra Papacal-Chuburná Puerto Km 5, Sierra Papacal, Mérida C.P. 97302, Yucatán, Mexico
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(23), 4499; https://doi.org/10.3390/polym15234499
Submission received: 13 October 2023 / Revised: 13 November 2023 / Accepted: 16 November 2023 / Published: 23 November 2023

Abstract

:
Most commercially available polymers are synthesized from compounds derived from petroleum, a finite resource. Because of this, there is a growing interest in the synthesis of new polymeric materials using renewable monomers. Following this concept, this work reports on the use of muconic acid as a renewable source for the development of new polyamides that can be used as proton-exchange membranes. Muconic acid was used as a comonomer in polycondensation reactions with 4,4′-(hexafluoroisopropylidene)bis(p-phenyleneoxy)dianiline, 2,5-diaminobencensulfonic acid, and 4,4′-diamino-2,2′-stilbenedisulfonic acid as comonomers in the synthesis of two new series of partially renewable aromatic–aliphatic polyamides, in which the degree of sulfonation was varied. Fourier transform infrared spectroscopy (FTIR) and nuclear magnetic resonance (1H, 13C, and 19F-NMR) techniques were used to confirm the chemical structures of the new polyamides. It was also observed that the degree of sulfonation was proportional to the molar ratio of the diamines in the feed. Subsequently, membranes were prepared by casting, and a complete characterization was conducted to determine their decomposition temperature (Td), glass transition temperature (Tg), density (ρ), and other physical properties. In addition, water uptake (Wu), ion-exchange capacity (IEC), and proton conductivity (σp) were determined for these membranes. Electrochemical impedance spectroscopy (EIS) was used to determine the conductivity of the membranes. MUFASA34 exhibited a σp value equal to 9.89 mS·cm−1, being the highest conductivity of all the membranes synthesized in this study.

Graphical Abstract

1. Introduction

The switch to renewable energy and the use of renewable resources is challenging; thus, many industrial processes are designed to work with fossil resources. In addition, in the past, the polymer industry designed materials with the only purpose of meeting the needs and optimizing production costs. This led to several issues, such as environmental pollution and the excessive use of non-renewable resources, among others. Consequently, polymers that can be more easily recycled and that bear some renewable moieties in their structures are currently being investigated [1,2,3].
An alternative technology for clean energy production is the proton-exchange membrane fuel cell (PEMFCs); this device exhibits the qualities of a high power density, easy start-up, scalable size, and low temperature operation; its function is to convert the chemical energy of fuels (such as H2, methanol, among others) directly into electrical energy using a proton-exchange membrane (PEM). The desired characteristics of PEMs are usually high proton conductivity; and thermal, chemical, and mechanical stability properties, not allowing the flow of reagents through the membrane and being economically accessible [4,5,6,7].
Accordingly, the PEM is one of the most important parts or PEMFCs. Nafion®, marketed by DuPont, is a sulfonated perfluorinated membrane used par excellence in PEMFCs; there are other similar membranes marketed by Dow and Asahi that have also been used as PEMs, however the cost of the membranes prepared from these ionomers is high due to the complicated production process, which limits the wide commercialization of these devices [7,8]. Due to the latter issue, two approaches can be considered to develop materials that can be used as PEMs: the first one is modifying the nafion to improve its properties, but this does not lower the cost of the membrane; the second one is synthesizing polymers with similar chemical structures. It is worth noting that significant progress in fuel cell development has also been made for aromatic polymer membranes [7,9,10,11,12,13] and acid-base blended proton-exchange membranes [14,15,16]. However, at present, almost all commercial synthetic polymers are derived from a small collection of petrochemical feedstocks [17].
In order to circumvent this issue, biomass is considered a very promising alternative to some compounds derived from fossil resources, and several polymers derived from renewable resources are currently being commercialized [18,19,20,21]. When superior mechanical and thermal properties are required, engineering polymers, such as polyamides (PAs), are preferred [22]; in this regard, some polyamides, which are synthesized from renewable raw materials, compete with their petroleum-derived counterparts from an economic perspective [17,22,23].
Muconic acid is a sustainable monomer obtained by biological routes, such as the biofermentation of sugar, lignin, or derivatives [24,25,26,27,28,29]. This compound is a high value-added product because it can be polymerized by different routes. The double bonds present in its structure make it possible to polymerize it by free radicals, while the presence of the two carboxylic acid groups allows polymerization by condensation in combination with diols or diamines [30]. In both types of polymerizations, the polymer backbone presents unsaturations, making it possible for these renewable-based materials to possess improved functionality in terms of their tunable chemical and mechanical properties [24,25,30,31].
Based on the abovementioned ideas, this research explores the synthesis and characterization of new sulfonated polyamides using muconic acid as a renewable comonomer in polycondensation reactions with 4,4′-(hexafluoroisopropylidene)bis(p-phenyleneoxy)dianiline (HFDA) and sulfonated diamine 2,5-diaminobencensulfonic acid (DABS) and 4,4′-diamino-2,2′-stilbenedisulfonic acid (DASDA) to produce two series of novel polyamides with a controlled degree of sulfonation, MUFABA and MUFASA, respectively. Their chemical structures are characterized by Fourier transform infrared (FTIR) and nuclear magnetic resonance (1H, 13C, 19F-NMR) spectroscopies. Electrochemical impedance spectroscopy (EIS) is used to determine the proton conductivity of the membrane of new muconic acid-based polyamide membranes. To the best of our knowledge, the use of muconic acid to obtain sulfonated polyamides for proton-exchange membranes has not been investigated. Therefore, this research contributes to the field of green chemistry by providing a sustainable approach to deliver specialty polymer products for alternative energy technologies.

2. Materials and Methods

2.1. Characterization Techniques

A Bruker Avance III HD (Bruker, Hamburg, Germany) was employed to record 1H-NMR, 13C-NMR, and 19F-NMR spectra at 400, 100, and 376 MHz, respectively. DMSO-d6 was used as a solvent at a typical concentration of 0.1 g·mL−1; hexafluorobenzene (HFB) and tetramethylsilane (TMS) were used as internal standards in this analysis. A Thermo Scientific Nicolet iS10 FTIR spectrometer fitted with an ATR accessory with a diamond crystal was employed to collect 32 spectra and coded for each membrane at 1 × 1 cm and around 400 μm in thickness in a range of 4000 to 650 cm−1 at a spectral resolution equal to 4 cm−1.
ATGA-DSC equipment (STA449F3 Jupiter; Netzsch, Selb, Germany) was used to determine the thermal stability of the films. For this purpose, the polymeric membrane was placed into an aluminum crucible, which was then placed in a TGA chamber. Subsequently, the sample was treated under a nitrogen flow and heated up from an ambient temperature to 600 °C at 10 °C·min−1. The decomposition temperature of the polymers was determined in a TA Instruments Thermogravimetric Analyzer TGA5500 (TA Instruments, New Castle, DE, USA), using samples of around 10 mg, at a temperature range of 30 °C to 800 °C, in a nitrogen atmosphere and a heating rate of 10 °C·min−1. A second-generation Bruker D2-Phaser diffractometer was employed to obtain the diffractograms corresponding to polymer films of 2 × 2 cm in size and around 400 μm in thickness. For this purpose, CuKα radiation (1.54 Å) was used at 30 kV, 10 mA, and a 2θ scale from 7 to 70°. A Sartorius analytical balance model, Quintix 124-1 s, was used to determine the density of the membranes through the flotation method using ethanol as the liquid, in ambient conditions. The density measurements were repeated five times and the average of these values was reported for each sample.
Elemental distribution scanning, from the polymers in a membrane form, was performed by Scanning Electron Microscopy using JEOL IT300 equipment (JEOL, Tokyo, Japan). Atomic force microscopy was measured with a Bruker model Multimode 8, using the tapping mode of a Sharp Nitride Lever probe SNL-10 (Bruker, CA, USA). Membranes of 0.5 × 0.5 cm and around 400 μm in thickness were used for SEM and AFM analyses. The Cannon-Ubbelohde viscometer No. 50 was used to determine the inherent viscosity, ηinh, of the polymers. Dissolutions of 0.2 g of each polymer in 100 mL of DMSO were prepared at 30 °C. The ηinh measurements were repeated five times under the given conditions and the average of the values obtained for each polymer sample was reported.
The proton conductivity, σp, was determined from EIS measurements using a Swagelok cell with 2 current collectors and electrodes of SS 316 using a Biologic VSP potentiostat with FRA, model VMP3B-10, from 1 MHz to 1 Hz at a 50 mV amplitude and open circuit potential under 100% relative humidity at 30 °C; the values were obtained from a Nyquist plot. The mathematical expression used for calculating the proton conductivity is as follows:
σ p = l m 10 × A T × R e
where σp (mS·cm−1) represents the proton conductivity, lm is the membrane thickness (µm), AT is the cross-section area of the membrane (cm2), and Re is the resistance value of the membrane obtained through the AC impedance method (Ω).
All characterizations were conducted in ambient conditions of humidity and temperature (55% RH and 25 °C, respectively) unless otherwise indicated.

2.2. Reagents

Comonomers: muconic acid (MUA), 4,4′-(hexafluoroisopropylidene)bis(p-phenyleneoxy)dianiline (HFDA), 2,5-diaminobencensulfonic acid (DABS), and 4,4′-Diamino-2,2′-stilbenedisulfonic acid (DASDA) were used as received. Calcium chloride (CaCl2), 1-methyl-2-pyrrolidinone (NMP), triphenylphosphite (TPP), and pyridine (Py) were used in the polycondensation reaction without further purification. Finally, solvents dimethyl sulfoxide (DMSO), dimethylformamide (DMF), and methanol were used as received. All chemical reactants and solvents were acquired from Sigma-Aldrich, Inc.

2.3. Synthesis of the Polyamides

A typical polymerization experiment was conducted as follows: an equimolar mixture of muconic acid and diamines were added to a 50 mL 3-neck flask equipped with a mechanical stirrer in a dry nitrogen atmosphere and according to the amounts shown in Table 1. Then, 2 mL of 1-methyl-2-pyrrolidinone (NMP) and 15 wt% calcium chloride (CaCl2) were also added and continuously stirred for 5 min. Then, 0.41 mL of pyridine and 0.41 mL of triphenylphosphite (TPP) were added and maintained with moderate stirring until a fully incorporated mixture was achieved (see Scheme 1 and Scheme 2). Finally, the mixture was heated at 110 °C for 12 h and kept under a constant-stirring condition. Then, the reaction mixture was cooled to room temperature and then precipitated into methanol. The polymer obtained was washed repeatedly with hot water to be purified, and finally it was dried at 100 °C in a vacuum oven for 24 h.

2.3.1. Characterization of Polymer MUFA

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.39 (–NH–, 2H), 7.77–7.75 (aromatic, 4H), 7.34–7.32 (aromatic, 4H), 7.13–7.04 (8H), 6.78 (2H), 6.48–6.23 (2H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.2. Characterization of Polymer MUFABA14

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.37 (–NH–, 8H), 8.36 (aromatic, 1H), 7.99 (aromatic, 2H), 7.77–7.75 (aromatic, 12H), 7.34–7.32 (aromatic, 12H), 7.13–7.04 (aromatic, 24H), 6.76 (8H), 6.48–6.22 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.3. Characterization of Polymer MUFABA24

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.68–10.38 (–NH–, 8H), 8.39 (aromatic, 2H), 7.99 (aromatic, 4H), 7.77–7.75 (aromatic, 8H), 7.35–7.33 (aromatic, 8H), 7.14–7.05 (aromatic, 16H), 6.76 (8H), 6.49–6.23 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.4. Characterization of Polymer MUFABA34

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.68–10.38 (–NH–, 8H), 8.37 (aromatic, 3H), 8.03 (aromatic, 6H), 7.77 (aromatic, 4H), 7.35–7.33 (aromatic, 4H), 7.14–7.05 (8H), 6.76 (8H), 6.48–6.23 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.5. Characterization of Polymer MUFASA14

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.37 (–NH–, 8H), 8.09 (aromatic, 2H), 8.05 (aromatic, 2H), 7.77–7.75 (aromatic, 12H), 7.64 (aromatic, 2H), 7.46 (CH=CH, 2H), 7.34–7.32 (aromatic, 12H), 7.13–7.04 (aromatic, 24H), 6.76 (8H), 6.48–6.23 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.6. Characterization of Polymer MUFASA24

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.49–10.39 (–NH–, 8H), 8.11 (aromatic, 4H), 8.02 (aromatic, 4H), 7.76 (aromatic, 8H), 7.64 (aromatic, 4H), 7.46 (CH=CH, 4H), 7.34–7.32 (aromatic, 8H), 7.13–7.05 (aromatic, 16H), 6.74 (8H), 6.48–6.23 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.3.7. Characterization of Polymer MUFASA34

1H-NMR (400 MHz, DMSO-d6, ppm): δ 10.50–10.40 (–NH–, 8H), 8.11 (aromatic, 6H), 8.01 (aromatic, 6H), 7.77–7.75 (aromatic, 4H), 7.66–7.62 (aromatic, 6H), 7.45 (CH=CH, 6H), 7.34–7.32 (aromatic, 4H), 7.12–7.04 (8H), 6.76 (8H), 6.48–6.23 (8H).
19F-NMR (376 MHz, DMSO-d6, ppm): δ −65.9.

2.4. Membrane Preparation, Ion-Exchange Capacity, and Water Uptake

Membranes were cast from polymeric DMSO solutions at 60 °C. The solution was filtered and poured onto a glass plate, and the solvent was slowly evaporated in a controlled DMSO atmosphere. Then, the membranes were subjected to treatment as described in the literature [32]. This treatment consisted firstly in the removal of the residual solvent by using methanol and deionized water, followed by activation with 1.0 N of hydrochloric acid. Finally, the membranes were dried under vacuum conditions at 150 °C for 24 h. The average thickness of the films was around 400 μm.
The experimental ion-exchange capacity of the polymer membrane was assessed, as described in the literature [32], by the titration method using the following mathematical expression:
I E C = V × M W d r y
where IEC (meq·g−1) represents the ion-exchange capacity, V (mL) represents the volume of the NaOH solution used in the titration, M is the molarity of the solution, and Wdry (g) is the mass of the dried membrane.
The theoretical ion-exchange capacity was calculated considering the complete incorporation of monomers and according to the following mathematical expression:
I E C T h e o = n M S × 1000 × s g w M U A + w H F D A + w M S
where IECTheo (meq·g−1) represents the theoretical ion-exchange capacity, nMS (mol) represents the moles of the sulfonated diamine present on the polymer, sg is the number of sulfonic groups corresponding to the DABS or DASDA sulfonated monomers, wMUA (g) is the weight of monomer MUA, wHFDA (g) is the weight of fluorinated monomer HFDA, and wMS (g) is the weight of sulfonated monomers DABS or DASDA.
The water uptake, Wu, of the polymer membrane was estimated, as described in the literature [33], by gravimetric measurements employing the mathematical expression given below:
W u = W w e t W d r y W d r y × 100
where Wu (%) represents the water uptake, Wwet (g) is the mass of the hydrated membrane, and Wdry (g) is the mass of the dried membrane.

3. Results and Discussion

Partially renewable muconic acid-based polyamides with an increasing degree of sulfonation were successfully prepared by a polycondensation reaction employing muconic acid as a renewable monomer and 4,4′-(hexafluoroisopropylidene)bis(p-phenyleneoxy)dianiline (HFDA) as a fluorinated comonomer, as well as 2,5-diaminobencensulfonic acid (DABS) and 4,4′-diamino-2,2′-stilbenedisulfonic acid (DASDA) as sulfonated comonomers (see Scheme 1 and Scheme 2, respectively). The roles of DABS and DASDA in this study were to incorporate sulfonic acid groups into the polymer backbone, thus endowing the resulting polymer with proton-exchange properties suitable for being used as a PEM. In addition, the use of both diamines allowed us to elucidate the effect of the number of sulfonic acid groups per repeating unit on the overall property balance.
It could be seen that the degree of sulfonation, DS, was effectively adapted by adjusting the molar ratio of diamines fed at the beginning of the reaction. Images of the raw DABS- and DASDA-containing partially renewable polyamides synthesized in this study are presented in Figure 1 and Figure 2, respectively. The non-sulfonated polyamide MUFA yielded yellow fibers and the sulfonated polyamide MUFABA series (Figure 1b–d) afforded fibers with coloration ranging from banana yellow to pineapple yellow, according to the increment in the sulfonation degree. In the same way, the sulfonated MUFASA series (Figure 2b–d) afforded fibers with coloration ranging from banana yellow to golden ochre. From these figures, it also can be seen that the raw material size decreases as the DS increases. According to the latter, the particle size is larger for the DABS-containing polyamides than for the MUFASA polyamides obtained from the same feed ratio. Polymeric membranes were prepared by casting and dissolving the polymers in DMSO. The membrane MUFA, Figure 1d, is transparent in appearance and quite resistant when touched; when the feed amount of sulfonated diamine increases in the polyamides, MUFABA or MUFASA series, the membranes are opaque with an orange-brown color that progressively intensifies as the DS increases.
The chemical structures of the new muconic acid-based polyamides were first confirmed by FTIR spectroscopy. The FTIR spectra of these new polyamides are shown in Figure 3; on the left side, the FTIR spectra of the series of polyamides containing DABS can be observed, while on the right side, those corresponding to the series of polyamides containing DASDA are shown. In both series, the characteristic absorption band for the N–H bond is observed at around 3300 cm−1. The band associated with the amide carbonyl groups (–CONH–) is displayed close to 1660 cm−1. The absorption band of C=C is exhibited at around 1500 cm−1. The band attributed to the C–F bond can be observed at around 1200 cm−1; this signal is more intense in the MUFA polymer, and it decreases as the concentration of the sulfonic acid increases. The presence of the sulfonic acid groups in the polymer backbone is indicated by the absorption bands shown at around 1086 and 1020 cm−1, which correspond to the asymmetric and symmetric O=S=O stretching vibrations of the –SO3H groups. It should be noted that these signals increase as the concentration of these sulfonic groups increase. The signal observed at around 3400 cm−1 can be attributed to the moisture present in the samples, which correlates with a higher concentration of sulfonic groups that makes these polymers more hydrophilic.
NMR spectroscopy was performed to confirm the chemical structures and compositions of the novel muconic acid-based polyamides, thus corroborating the results obtain by FTIR. The 1H-NMR spectra of the MUFABA and MUFASA polyamide series are shown in Figure 4 and Figure 5, respectively. This analysis indicated that the ratio of the proton integration areas agreed quite well with the expected chemical structures of the polyamides. For instance, in Figure 4, a single signal ascribed to the proton attached to the amide group (Hc) in the polyamide MUFA can be observed at δ = 10.39 ppm. For the DABS-containing polyamide series, the proton attached to the amide group displays two signals, in δ ≈ 10.68 (Hg) and 10.38 ppm (Hc), and it can also be seen that the new signal at 10.68 ppm increases as the concentration of DABS in the polymer increases. Likewise, the signals observed in the region of 8.4–7.9 ppm are attributed to the aromatic protons in the DABS moiety (Hh, Hi, Hj). The signals attributed to the incorporation of the HFDA monomer appear at around δ ≈ 7.7 (Hf), 7.3 (Hd), and 7.11–7.04 ppm (He), while the olefinic proton signals ascribed to the incorporation of the muconic acid monomer (Ha, Hb) are observed at around 6.8–6.7 ppm; these monomer-derived signals are common for all polyamides synthesized in this study. Finally, in Figure 5, the DASDA-containing polyamide series presents multiple signals attributed to the proton attached to the amide group, in δ ≈ 10.50 (Hg, Hl) and 10.40 ppm (Hc). Moreover, the signals observed in the region of 8.1–7.6 ppm are attributed to the aromatic protons in the DASDA moiety (Hh, Hi, Hj, Hk).
The intensity of these signals varied progressively according to the fluorinated/sulfonated diamine feed molar ratio. The triplet peaks at around δ ≈ 8.5 ppm in the 1H-NMR spectra were attributed to pyridine, a reactive used in the synthesis of these polyamides, which was not completely removed after the purification procedure. The ratio between the proton integration areas that remained unchanged in the polyamide MUFA and the series of sulfonated polyamides, MUFABA and MUFASA, were employed to calculate the DS. The values of DS, determined from the 1H-NMR spectra, for the muconic acid-based polyamides MUFABA14; MUFABA24; and MUFABA34 were found to be 24.3%; 48.7%; and 72.2%, respectively, while those of the polyamides MUFASA14; MUFASA24; and MUFASA34 were found to be 22.8%; 46.1%; and 70.6%, respectively. These DS values were close and could be correlated to the diamine ratio established in the feed values (HFDA, and DABS or DASDA) of the new polyamides, which were 25%, 50%, and 75% with respect to the sulfonated diamine, for each series, respectively (see Table 1). These values indicated that the incorporation of sulfonic groups was effectively controlled.
The Tg of the polymers was determined by DSC, and their thermograms are shown in Figure 6; those corresponding to the DABS-containing polyamide series are displayed on the left side, while those corresponding to the DASDA-containing polyamide series are shown on the right. The DSC curves present some features lower than 200 °C; those observed around 100 °C are attributed to the humidity adsorbed by the membranes, while those observed near 200 °C are attributed to the residual solvent (DMSO) used in the membrane casting. Therefore, the glass transition temperature of the polymer was considered to be the first characteristic above 200 °C. It can be seen that the Tg for the non-sulfonated polymer MUFA is about 241 °C, a high Tg that is ascribed to the rigidity of the main chain caused by the aromatic rings of the HFDA monomer. It was found that for the MUFABA series, the Tg remained almost the same. The Tg values for MUFABA14, MUFABA24, and MUFABA34 were found to be 244 °C, 240 °C, and 241 °C, respectively. The latter can be attributed to two opposite effects. On the one hand, as DS increases, the ratio of fluorinated/sulfonated diamine decreases, and, in turn, so does the rigidity of the polymer chains. On the other hand, the incorporation of sulfonic acid groups creates strong ionic interactions between polymeric chains, increasing their rigidity. This contrasting behavior results in a negligible change in Tg as the quantity of the sulfonated moiety is varied in the polymer. It was also found that for the MUFASA series, the Tg increased as the DS increased. The Tg values for MUFASA14, MUFASA24, and MUFASA34 were found to be 217 °C, 229 °C, and 232 °C, respectively. This increase in the Tg value of the MUFASA series could be due to the higher concentration of sulfonic groups in the polymer backbone, which inhibited the relaxation process of polymer main chains as a result of strong ionic interactions between them.
The thermal stability of polyamides was studied by TGA under a nitrogen atmosphere. The thermograms are presented in Figure 7 with a 2% offset between each thermogram. It can be observed that the polymer MUFA exhibits a decomposition temperature, Td, of about 322 °C. For the MUFABA series, the Td values observed were 318 °C, 303 °C, and 273 °C for MUFABA14, MUFABA24, and MUFABA34, respectively, while for the MUFASA series, the Td values observed were 287 °C, 272 °C, and 276 °C for MUFASA14, MUFASA24, and MUFASA34, respectively. From these results, it is clear that the introduction of labile groups, such as –SO3H, into the macromolecular architecture is reflected by a lower Td value.
The mean intersegmental distance or d-spacing between the polymer chains was calculated by the Bragg’s equation, d i = n λ 2 s i n θ ; the angle was determined at the maximum intensity of the amorphous peak. In general, the XRD technique showed patterns characteristic of amorphous polymeric materials, and the absences of narrow peaks showed that these samples did not exhibit any degree of crystallinity, with the exception of the MUFASA34 sample, in which the highest quantity of the DASDA monomer incorporated seemed to induce a certain degree of crystallinity in the resulting polymer (see Figure 8). The maximum reflective intensity was observed at interval of 2θ ≈ 17.5–24.6°. This angle increased as the degree of sulfonation increased.
The density, ρ, of the new polyamide membranes was determined by the flotation method using ethanol as the liquid, at room temperature. The values, shown in Table 2, were found to range from 1.38 to 1.45 g·cm−3. The inherent viscosity (ηinh), also shown in Table 2, presents values ranging from 0.352 to 0.201 dL·g−1. It can be seen that the lowest inherent viscosity is that of the MUFASA34 polymer, indicating that the chain size of this polymer is smaller than that of the other polymers. The ηinh values obtained are in the range of other synthetic polymers [34,35]. The ηinh values of the DABS-containing polyamides are very similar to each other, whereas the ηinh values of the DASDA-containing polyamides decrease as the DS increases; it is likely that the higher amount of DASDA in the feed causes the insolubility of the reactants in the polymer solution, affecting the growth of the polymer.
The morphology of the membranes was observed by AFM and the topographic images obtained by this technique, in tapping mode, are presented in Figure 9. The morphology of the MUFABA14 sample is mainly characterized by voids and a few areas with protuberances that can be associated with hydrophobic fluorine-rich regions. The MUFABA24 sample presents a homogeneous surface morphology, in which the existence of voids or large protuberances is not appreciated; this can be attributed to the balance between the hydrophobic and hydrophilic regions created by the sulfonic acid groups. In contrast, the MUFABA34 sample shows a quite irregular surface morphology consisting of plateaus immersed in regions with protuberances of different sizes that can be attributed to the higher number of –SO3H groups in this entire series of polyamides, which causes the disruption of the chain-packing process. The topology of MUFASA14 is mainly characterized by a regular surface formed by small protuberances that lacks voids and resembles that of MUFABA14. The MUFASA24 sample exhibits a surface morphology that is composed of different plateaus and some areas with pinnacles in which phase segregation can be noticed. Finally, the MUFASA34 sample shows a surface on which quite large protuberances can be seen. Due to the high DS of this polyamide, it is likely that large hydrophilic regions are formed and these begin to segregate from the hydrophobic regions, giving rise to large clusters. These observations indicate the existence of phase separation, which is a fundamental aspect to achieve high proton conductivity, since the segregation of hydrophilic and hydrophobic phases builds the channels through which ionic transport is conducted.
Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS) were employed to study the surfaces of the polyamide samples with the highest content of sulfonic acid groups in each series. Representative results are shown in Figure 10 and Figure 11, corresponding to polymers MUFABA34 and MUFASA34, respectively. The SEM images show the heterogeneous surface texture and morphology for these partially renewable ionomers that can be attributed to phase segregation. Furthermore, EDS was employed to estimate the chemical composition of the polymers. As can be seen, both polyamide samples mainly show the presence of carbon (C), oxygen (O), sulfur (S), fluorine (F), nitrogen (N), and calcium (Ca). These results, specifically the sulfur weight percent and fluorine weight percent determined by EDS, are in agreement with those expected from the fluorinated/sulfonated diamine feed molar ratio.
The electrochemical impedance spectroscopy (EIS) measurements of the sulfonated membranes were performed under 100% relative humidity at 30 °C in the range of 1 MHz to 1 Hz; Nyquist plots of the most sulfonated membranes of each series, MUFABA34 and MUFASA34, are shown in Figure 12. The resistance, Re, of the membranes was calculated using the “ZFit” function of EC-Lab software V11.27 from the intersection with the abscissa axis, which corresponds to the resistance value, as can be observed in the insets of Figure 12; all measures are repeated three times in the given conditions, and the averages of the results obtained are reported. Then, the proton conductivity, σp, was calculated by Equation (1), and the results are presented in Table 3. It can be observed that the σp value increases as the sulfonic acid group concentration increases. The σp values for the MUFABA polyamide series were found to range from 0.027 to 1.608 mS·cm−1, while those of the MUFASA polyamide series were found to range from 0.367 to 9.895 mS·cm−1. It is important to mention that, when comparing a polyamide from the MUFABA series with one from the MUFASA series with similar DS values, the latter actually contains twice as many –SO3H groups as the former, because the diamine DASDA, which gives rise to the MUFASA series, contains one more –SO3H group than the diamine DABS, which gives rise to the MUFABA series. For this reason, the MUFASA34 membrane, with a DS = 70.6%, shows higher IEC and σ p values than the MUFABA34 membrane that shows a DS = 72.2%. The polyamide MUFASA34 is the one that contains the highest number of sulfonic acid groups of both series of new polyamides. Due to the latter reason, it also shows the highest IEC and Wu values, which in turn is reflected in the highest proton conductivity of all the materials synthesized in this study.
Various efforts have been made to develop sulfonated polymeric materials with a sustainable approach around the world [36,37,38,39,40]. Table 4 shows some of these materials, along with their ionic properties, and compares them with the values obtained for the muconic acid-based polymers developed in this research. The results obtained for the sulfonated polymers MUFABA34 and MUFASA34 are comparable with other fully or partially renewable polymers bearing sulfonic acid groups. The latter indicates that the sustainable methodology addressed in this study is an effective tool to offer special environmentally friendly polymers.

4. Conclusions

The muconic acid monomer reacted successfully in polycondensation reactions with two aromatic-sulfonated diamines and a fluorinated diamine as comonomers for the synthesis of two series of partially renewable aromatic–aliphatic polyamides with an increasing degree of sulfonation (DS). FTIR and NMR spectroscopy methods were employed to confirm the polymer chemical structures and also revealed that the DS was effectively tailored by adjusting the feed molar ratio of the diamines. It was also found that the Wu, IEC, and σp values increased together with the DS. The highest value of σp determined by electrochemical impedance spectroscopy (EIS) was found to be 9.895 mS·cm−1 at 30 °C corresponding to the MUFASA34 polyamide. The ionic properties reported in this study for the partially renewable muconic acid-based ionomers were close to other fully or partially renewable sulfonated polymers. It is important to note that the incorporation of renewable monomers is a powerful tool to synthesize sustainable and eco-friendly specialty polymers that can be an alternative for PEM applications.

Author Contributions

Conceptualization, A.A.S. and J.V.; data curation, C.C.-G., S.R.V.-G. and A.A.S.; investigation, C.C.-G. and A.O.; methodology, C.C.-G., T.E.S. and D.E.P.-C.; writing—original draft, C.C.-G. and J.V.; writing—review and editing, C.C.-G. and J.V. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from DGAPA-UNAM PAPIIT through project IN108022 is gratefully acknowledged.

Data Availability Statement

The data presented in this study are available from the corresponding authors upon reasonable request.

Acknowledgments

We are grateful to Gerardo Cedillo Valverde, Karla Eriseth Reyes Morales, M.C. José Martín Baas López, and M.C. Enrique Escobedo Hernández for their assistance in the NMR, thermal properties, EIS, and AFM experiments, respectively. Financial support from the National Council for Science and Technology of Mexico (CONACyT) (PhD Scholarship given to C.C-G.; number 762228) is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Park, M.; Hong, S.-J.; Kim, N.-K.; Shin, J.; Kim, Y.-W. Vegetable Oil-Derived Polyamide Multiblock Copolymers toward Chemically Recyclable Pressure-Sensitive Adhesives. ACS Sustain. Chem. Eng. 2023, 11, 10095–10107. [Google Scholar] [CrossRef]
  2. Bottega Pergher, B.; Girigan, N.; Vlasblom, S.; Weinland, D.H.; Wang, B.; van Putten, R.-J.; Gruter, G.-J.M. Reactive Phenolic Solvents Applied to the Synthesis of Renewable Aromatic Polyesters with High Isosorbide Content. Polym. Chem. 2023, 14, 3225–3238. [Google Scholar] [CrossRef]
  3. Mu, T.; Leng, S.; Tang, W.; Shi, N.; Wang, G.; Yang, J. High-Performance and Low-Cost Membranes Based on Poly(Vinylpyrrolidone) and Cardo-Poly(Etherketone) Blends for Vanadium Redox Flow Battery Applications. Batteries 2022, 8, 230. [Google Scholar] [CrossRef]
  4. Mukoma, P.; Jooste, B.R.; Vosloo, H.C.M. Synthesis and Characterization of Cross-Linked Chitosan Membranes for Application as Alternative Proton Exchange Membrane Materials in Fuel Cells. J. Power Sources 2004, 136, 16–23. [Google Scholar] [CrossRef]
  5. Li, N.; Shin, D.W.; Hwang, D.S.; Lee, Y.M.; Guiver, M.D. Polymer Electrolyte Membranes Derived from New Sulfone Monomers with Pendent Sulfonic Acid Groups. Macromolecules 2010, 43, 9810–9820. [Google Scholar] [CrossRef]
  6. Li, J.; Pan, M.; Tang, H. Understanding Short-Side-Chain Perfluorinated Sulfonic Acid and Its Application for High Temperature Polymer Electrolyte Membrane Fuel Cells. RSC Adv. 2014, 4, 3944–3965. [Google Scholar] [CrossRef]
  7. Iulianelli, A.; Clarizia, G.; Gugliuzza, A.; Ebrasu, D.; Bevilacqua, A.; Trotta, F.; Basile, A. Sulfonation of PEEK-WC Polymer via Chloro-Sulfonic Acid for Potential PEM Fuel Cell Applications. Int. J. Hydrogen Energy 2010, 35, 12688–12695. [Google Scholar] [CrossRef]
  8. Ahmad, M.I.; Zaidi, S.M.J.; Rahman, S.U. Proton Conductivity and Characterization of Novel Composite Membranes for Medium-Temperature Fuel Cells. Desalination 2006, 193, 387–397. [Google Scholar] [CrossRef]
  9. Xing, P.; Robertson, G.P.; Guiver, M.D.; Mikhailenko, S.D.; Wang, K.; Kaliaguine, S. Synthesis and Characterization of Sulfonated Poly(Ether Ether Ketone) for Proton Exchange Membranes. J. Memb. Sci. 2004, 229, 95–106. [Google Scholar] [CrossRef]
  10. Devrim, Y.; Erkan, S.; Baç, N.; Eroǧlu, I. Preparation and Characterization of Sulfonated Polysulfone/Titanium Dioxide Composite Membranes for Proton Exchange Membrane Fuel Cells. Int. J. Hydrogen Energy 2009, 34, 3467–3475. [Google Scholar] [CrossRef]
  11. Santiago, A.A.; Vargas, J.; Fomine, S.; Gaviño, R.; Tlenkopatchev, M.A. Polynorbornene with Pentafluorophenyl Imide Side Chain Groups: Synthesis and Sulfonation. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 2925–2933. [Google Scholar] [CrossRef]
  12. Zhang, Z.; Liu, H.; Dong, T.; Deng, Y.; Li, Y.; Lu, C.; Jia, W.; Meng, Z.; Zhou, M.; Tang, H. Phosphonate Poly(Vinylbenzyl Chloride)-Modified Sulfonated Poly(Aryl Ether Nitrile) for Blend Proton Exchange Membranes: Enhanced Mechanical and Electrochemical Properties. Polymers 2023, 15, 3203. [Google Scholar] [CrossRef]
  13. Zavorotnaya, U.M.; Ponomarev, I.I.; Volkova, Y.A.; Sinitsyn, V.V. Development of High-Performance Hydrogen-Air Fuel Cell with Flourine-Free Sulfonated Co-Polynaphthoyleneimide Membrane. Membranes 2023, 13, 485. [Google Scholar] [CrossRef] [PubMed]
  14. Kim, A.R.; Poudel, M.B.; Chu, J.Y.; Vinothkannan, M.; Santhosh Kumar, R.; Logeshwaran, N.; Park, B.H.; Han, M.K.; Yoo, D.J. Advanced Performance and Ultra-High, Long-Term Durability of Acid-Base Blended Membranes over 900 Hours Containing Sulfonated PEEK and Quaternized Poly(Arylene Ether Sulfone) in H2/O2 Fuel Cells. Compos. Part B Eng. 2023, 254, 110558. [Google Scholar] [CrossRef]
  15. Jiang, J.; Xiao, M.; Huang, S.; Han, D.; Wang, S.; Meng, Y. Phosphonic Acid-Imidazolium Containing Polymer Ionomeric Membranes Derived from Poly (Phenylene Oxide) towards Boosting the Performance of HT-PEM Fuel Cells. J. Memb. Sci. 2023, 686, 121982. [Google Scholar] [CrossRef]
  16. Li, G.; Shen, R.; Hu, S.; Wang, B.; Algadi, H.; Wang, C. Norbornene-Based Acid–Base Blended Polymer Membranes with Low Ion Exchange Capacity for Proton Exchange Membrane Fuel Cell. Adv. Compos. Hybrid Mater. 2022, 5, 2131–2137. [Google Scholar] [CrossRef]
  17. Woroch, C.P.; Cox, I.W.; Kanan, M.W. A Semicrystalline Furanic Polyamide Made from Renewable Feedstocks. J. Am. Chem. Soc. 2023, 145, 697–705. [Google Scholar] [CrossRef]
  18. Reinaldo, J.S.; Milfont, C.H.R.; Gomes, F.P.C.; Mattos, A.L.A.; Medeiros, F.G.M.; Lopes, P.F.N.; Filho, M.d.S.M.S.; Matsui, K.N.; Ito, E.N. Influence of Grape and Acerola Residues on the Antioxidant, Physicochemical and Mechanical Properties of Cassava Starch Biocomposites. Polym. Test. 2021, 93, 107015. [Google Scholar] [CrossRef]
  19. Zarna, C.; Opedal, M.T.; Echtermeyer, A.T.; Chinga-Carrasco, G. Reinforcement Ability of Lignocellulosic Components in Biocomposites and Their 3D Printed Applications–A Review. Compos. Part C Open Access 2021, 6, 100171. [Google Scholar] [CrossRef]
  20. Terroba-Delicado, E.; Fiori, S.; Gomez-Caturla, J.; Montanes, N.; Sanchez-Nacher, L.; Torres-Giner, S. Valorization of Liquor Waste Derived Spent Coffee Grains for the Development of Injection-Molded Polylactide Pieces of Interest as Disposable Food Packaging and Serving Materials. Foods 2022, 11, 1162. [Google Scholar] [CrossRef] [PubMed]
  21. Borrero-l, A.M.; Valencia, C.; Franco, J.M. Lignocellulosic Materials for the Production of Biofuels, Biochemicals and Biomaterials and Applications Of. Polymers 2022, 14, 881. [Google Scholar] [CrossRef]
  22. Wolff, S.; Rüppel, A.; Rida, H.A.; Heim, H.-P. Emission and Mechanical Properties of Glass and Cellulose Fiber Reinforced Bio-Polyamide Composites. Polymers 2023, 15, 2603. [Google Scholar] [CrossRef]
  23. Beppu, S.; Tachibana, Y.; Kasuya, K.I. Recyclable Polycarbosilane from a Biomass-Derived Bifuran-Based Monomer. ACS Macro Lett. 2023, 12, 536–542. [Google Scholar] [CrossRef]
  24. Khalil, I.; Quintens, G.; Junkers, T.; Dusselier, M. Muconic Acid Isomers as Platform Chemicals and Monomers in the Biobased Economy. Green Chem. 2020, 22, 1517–1541. [Google Scholar] [CrossRef]
  25. Quintens, G.; Vrijsen, J.H.; Adriaensens, P.; Vanderzande, D.; Junkers, T. Muconic Acid Esters as Bio-Based Acrylate Mimics. Polym. Chem. 2019, 10, 5555–5563. [Google Scholar] [CrossRef]
  26. Liu, P.; Zheng, Y.; Yuan, Y.; Zhang, T.; Su, T.; Li, Q.; Liang, Q.; Qi, Q. A Circular Bioprocess for the Sustainable Conversion of Polyethylene Terephthalate to Muconic Acid with an Engineered Pseudomonas Putida. SSRN Electron. J. 2022. [Google Scholar] [CrossRef]
  27. Molinari, F.; Pollegioni, L.; Rosini, E. Whole-Cell Bioconversion of Renewable Biomasses-Related Aromatics to Cis,Cis-Muconic Acid. ACS Sustain. Chem. Eng. 2023, 11, 2476–2485. [Google Scholar] [CrossRef]
  28. Nandhini, R.; Sivaprakash, B.; Rajamohan, N.; Vo, D.V.N. Lignin and Polylactic Acid for the Production of Bioplastics and Valuable Chemicals. Environ. Chem. Lett. 2022, 21, 403–427. [Google Scholar] [CrossRef]
  29. Weiland, F.; Barton, N.; Kohlstedt, M.; Becker, J.; Wittmann, C. Systems Metabolic Engineering Upgrades Corynebacterium Glutamicum to High-Efficiency Cis, Cis-Muconic Acid Production from Lignin-Based Aromatics. Metab. Eng. 2023, 75, 153–169. [Google Scholar] [CrossRef] [PubMed]
  30. Maniar, D.; Fodor, C.; Karno Adi, I.; Woortman, A.J.J.; van Dijken, J.; Loos, K. Enzymatic Synthesis of Muconic Acid-Based Polymers: Trans, Trans-Dimethyl Muconate and Trans, β-Dimethyl Hydromuconate. Polymers 2021, 13, 2498. [Google Scholar] [CrossRef] [PubMed]
  31. Yan, K.; Wang, J.; Wang, Z.; Yuan, L. Bio-Based Monomers for Amide-Containing Sustainable Polymers. Chem. Commun. 2023, 59, 382–400. [Google Scholar] [CrossRef]
  32. Santiago, A.A.; Ibarra-Palos, A.; Cruz-Morales, J.A.; Sierra, J.M.; Abatal, M.; Alfonso, I.; Vargas, J. Synthesis, Characterization, and Heavy Metal Adsorption Properties of Sulfonated Aromatic Polyamides. High Perform. Polym. 2018, 30, 591–601. [Google Scholar] [CrossRef]
  33. Ruiz, I.; Corona-García, C.; Santiago, A.A.; Abatal, M.; Téllez Arias, M.G.; Alfonso, I.; Vargas, J. Synthesis, Characterization, and Assessment of Novel Sulfonated Polynorbornene Dicarboximides as Adsorbents for the Removal of Heavy Metals from Water. Environ. Sci. Pollut. Res. 2021, 28, 52014–52031. [Google Scholar] [CrossRef] [PubMed]
  34. Mehdipour-Ataei, S.; Hatami, M. Synthesis and Characterization of Novel Heat Resistant Poly(Amide Imide)S. Eur. Polym. J. 2005, 41, 2010–2015. [Google Scholar] [CrossRef]
  35. Pali-Casanova, R.d.J.; Yam-Cervantes, M.A.; Zavala-Loría, J.d.C.; Loría-Bastarrachea, M.I.; Aguilar-Vega, M.d.J.; Dzul-López, L.A.; Sámano-Celorio, M.L.; Crespo-álvarez, J.; García-Villena, E.; Agudo-Toyos, P.; et al. Effect of Sulfonic Groups Concentration on IEC Properties in New Fluorinated Copolyamides. Polymers 2019, 11, 1169. [Google Scholar] [CrossRef]
  36. Ngadiwiyana; Gunawan; Prasetya, N.B.A.; Kusworo, T.D.; Susanto, H. Synthesis and Characterization of Sulfonated Poly(Eugenol-Co-Allyleugenol) Membranes for Proton Exchange Membrane Fuel Cells. Heliyon 2022, 8, e12401. [Google Scholar] [CrossRef]
  37. Pawlicka, A.; Mattos, R.I.; Tambelli, C.E.; Silva, I.D.A.; Magon, C.J.; Donoso, J.P. Magnetic Resonance Study of Chitosan Bio-Membranes with Proton Conductivity Properties. J. Memb. Sci. 2013, 429, 190–196. [Google Scholar] [CrossRef]
  38. Corona-García, C.; Onchi, A.; Santiago, A.A.; Martínez, A.; Pacheco-Catalán, D.E.; Alfonso, I.; Vargas, J. Synthesis and Characterization of Partially Renewable Oleic Acid-Based Ionomers for Proton Exchange Membranes. Polymers 2021, 13, 130. [Google Scholar] [CrossRef]
  39. Pasini Cabello, S.D.; Ochoa, N.A.; Takara, E.A.; Mollá, S.; Compañ, V. Influence of Pectin as a Green Polymer Electrolyte on the Transport Properties of Chitosan-Pectin Membranes. Carbohydr. Polym. 2017, 157, 1759–1768. [Google Scholar] [CrossRef]
  40. Farzin, S.; Johnson, T.J.; Chatterjee, S.; Zamani, E.; Dishari, S.K. Ionomers From Kraft Lignin for Renewable Energy Applications. Front. Chem. 2020, 8, 562278. [Google Scholar] [CrossRef]
Scheme 1. Synthesis route of DABS-containing muconic acid-based polyamides.
Scheme 1. Synthesis route of DABS-containing muconic acid-based polyamides.
Polymers 15 04499 sch001
Scheme 2. Synthesis route of DASDA-containing muconic acid-based polyamides.
Scheme 2. Synthesis route of DASDA-containing muconic acid-based polyamides.
Polymers 15 04499 sch002
Figure 1. Photographic images of the synthesized raw muconic acid-based polyamides (top) and their corresponding polymeric membranes (bottom): (a) MUFA, (b) MUFABA14, (c) MUFABA24, and (d) MUFABA34.
Figure 1. Photographic images of the synthesized raw muconic acid-based polyamides (top) and their corresponding polymeric membranes (bottom): (a) MUFA, (b) MUFABA14, (c) MUFABA24, and (d) MUFABA34.
Polymers 15 04499 g001
Figure 2. Photographic images of the synthesized raw muconic acid-based polyamides (top) and their corresponding polymeric membranes (bottom): (a) MUFA, (b) MUFASA14, (c) MUFASA24, and (d) MUFASA34.
Figure 2. Photographic images of the synthesized raw muconic acid-based polyamides (top) and their corresponding polymeric membranes (bottom): (a) MUFA, (b) MUFASA14, (c) MUFASA24, and (d) MUFASA34.
Polymers 15 04499 g002
Figure 3. FTIR spectra of the muconic acid-based polyamides: (Left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, (d) MUFA; (Right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Figure 3. FTIR spectra of the muconic acid-based polyamides: (Left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, (d) MUFA; (Right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Polymers 15 04499 g003
Figure 4. 1H-NMR spectra of the muconic acid-based polyamides: (a) MUFABA14, (b) MUFABA24, (c) MUFABA34, and (d) MUFA. Lower case letters correspond to the protons indicated in the chemical structure.
Figure 4. 1H-NMR spectra of the muconic acid-based polyamides: (a) MUFABA14, (b) MUFABA24, (c) MUFABA34, and (d) MUFA. Lower case letters correspond to the protons indicated in the chemical structure.
Polymers 15 04499 g004
Figure 5. 1H-NMR spectra of the muconic acid-based polyamides: (a) MUFASA34, (b) MUFASA24, (c) MUFASA14, and (d) MUFA. Lower case letters correspond to the protons indicated in the chemical structure.
Figure 5. 1H-NMR spectra of the muconic acid-based polyamides: (a) MUFASA34, (b) MUFASA24, (c) MUFASA14, and (d) MUFA. Lower case letters correspond to the protons indicated in the chemical structure.
Polymers 15 04499 g005
Figure 6. DSC thermograms of the DABS-containing polyamide series (left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, and (d) MUFA, and the DASDA-containing polyamide series (right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Figure 6. DSC thermograms of the DABS-containing polyamide series (left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, and (d) MUFA, and the DASDA-containing polyamide series (right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Polymers 15 04499 g006
Figure 7. TGA thermograms of the DABS-containing polyamide series (left): (a) MUFA, (b) MUFABA14, (c) MUFABA24, and (d) MUFABA34, and the DASDA-containing polyamide series (right): (a) MUFA, (e) MUFASA14, (f) MUFASA24, and (g) MUFASA34.
Figure 7. TGA thermograms of the DABS-containing polyamide series (left): (a) MUFA, (b) MUFABA14, (c) MUFABA24, and (d) MUFABA34, and the DASDA-containing polyamide series (right): (a) MUFA, (e) MUFASA14, (f) MUFASA24, and (g) MUFASA34.
Polymers 15 04499 g007
Figure 8. XRD patterns of the DABS-containing polyamide series (left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, and (d) MUFA, and the DASDA-containing polyamide series (right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Figure 8. XRD patterns of the DABS-containing polyamide series (left): (a) MUFABA34, (b) MUFABA24, (c) MUFABA14, and (d) MUFA, and the DASDA-containing polyamide series (right): (e) MUFASA34, (f) MUFASA24, (g) MUFASA14, and (d) MUFA.
Polymers 15 04499 g008
Figure 9. Three-dimensional AFM micrographs (5 × 5 μm) of (a) MUFABA14, (b) MUFABA24, (c) MUFABA34, (d) MUFASA14, (e) MUFASA24, and (f) MUFASA34.
Figure 9. Three-dimensional AFM micrographs (5 × 5 μm) of (a) MUFABA14, (b) MUFABA24, (c) MUFABA34, (d) MUFASA14, (e) MUFASA24, and (f) MUFASA34.
Polymers 15 04499 g009
Figure 10. SEM image and EDS spectra of MUFABA34.
Figure 10. SEM image and EDS spectra of MUFABA34.
Polymers 15 04499 g010
Figure 11. SEM image and EDS spectra of MUFASA34.
Figure 11. SEM image and EDS spectra of MUFASA34.
Polymers 15 04499 g011
Figure 12. Nyquist plots of (a) MUFABA34 and (b) MUFASA34. Inset: zoomed-in view of the Nyquist plots in the high-frequency-region membrane.
Figure 12. Nyquist plots of (a) MUFABA34 and (b) MUFASA34. Inset: zoomed-in view of the Nyquist plots in the high-frequency-region membrane.
Polymers 15 04499 g012
Table 1. Molar relations of monomers in the feed for the synthesis of the muconic acid-based polyamide series.
Table 1. Molar relations of monomers in the feed for the synthesis of the muconic acid-based polyamide series.
PolymerMUA a
(mmol)
HFDA b
(mmol)
DABS c
(mmol)
DASDA d
(mmol)
MUFA0.640.640.000.00
MUFABA140.640.480.160.00
MUFABA240.640.320.320.00
MUFABA340.640.160.480.00
MUFASA140.640.480.000.16
MUFASA240.640.320.000.32
MUFASA340.640.160.000.48
a Muconic acid, b 4,4′-(hexafluoroisopropylidene)bis(p-phenyleneoxy)dianiline, c 2,5-diaminobencensulfonic acid, d 4,4′-Diamino-2,2′-stilbenedisulfonic acid.
Table 2. Physical properties of the muconic acid-based polyamide series.
Table 2. Physical properties of the muconic acid-based polyamide series.
PolymerTg
(°C) a
Td
(°C) b
di
(Å) c
ρ
(g·cm−3) d
ηinh
(dL·g−1) e
MUFA2413222.551.380.352
MUFABA142443182.051.380.343
MUFABA242403032.271.390.304
MUFABA342412731.851.430.348
MUFASA142172871.951.370.266
MUFASA242292721.971.380.213
MUFASA342322762.051.450.201
a Glass transition temperature determined by DSC. b Decomposition temperature determined by TGA. c Average separation distance between polymer chains determined by XRD. d Density determined by the flotation method. e Inherent viscosity at a polymer concentration of 0.2 g·dL−1.
Table 3. Ionic properties of novel muconic acid-based polyamides.
Table 3. Ionic properties of novel muconic acid-based polyamides.
PolymerDS a
(%)
IEC (mmol·g−1)Wu
(%)
σ p
(mS·cm−1c
TheoreticalExperimental b
MUFA---4.51-
MUFABA1424.30.430.377.870.027
MUFABA2448.71.010.9629.170.426
MUFABA3472.21.821.8033.491.608
MUFASA1422.80.800.7212.450.367
MUFASA2446.11.701.8028.780.849
MUFASA3470.62.732.8136.939.895
a Determinated by 1H-NMR. b Determined by titration. c Determinated by EIS.
Table 4. Ionic properties’ comparison of the muconic acid-based ionomers reported in this study and some renewable polymers reported in the literature.
Table 4. Ionic properties’ comparison of the muconic acid-based ionomers reported in this study and some renewable polymers reported in the literature.
PolymerIECExp
(meq·g−1) a
Wu
(%) b
σp
(mS·cm) c
Reference
Author
(SPEAE) d0.35651.20.018[36]
S5 e--0.005[37]
DASA f2.244.10.5[38]
CH g0.372492.44[39]
DAFASA2/4 f0.817.11.5[38]
LS 1.6 h1.625.98.12[40]
MUFABA341.8033.491.608This study
MUFASA342.8136.939.895This study
a Ion-exchange capacity determined by titration. b Water uptake determined by gravimetric measurements. c Proton conductivity determined by EIS. d Sulfonated poly(eugenolco-allyleugenol). e Chitosan-based membranes containing acetic acid (HAC) and plasticized with glycerol. f Oleic acid-based polyamide. g Pristine chitosan. h Ionomer from kraft lignin.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Corona-García, C.; Onchi, A.; Santiago, A.A.; Soto, T.E.; Vásquez-García, S.R.; Pacheco-Catalán, D.E.; Vargas, J. Synthesis, Characterization, and Proton Conductivity of Muconic Acid-Based Polyamides Bearing Sulfonated Moieties. Polymers 2023, 15, 4499. https://doi.org/10.3390/polym15234499

AMA Style

Corona-García C, Onchi A, Santiago AA, Soto TE, Vásquez-García SR, Pacheco-Catalán DE, Vargas J. Synthesis, Characterization, and Proton Conductivity of Muconic Acid-Based Polyamides Bearing Sulfonated Moieties. Polymers. 2023; 15(23):4499. https://doi.org/10.3390/polym15234499

Chicago/Turabian Style

Corona-García, Carlos, Alejandro Onchi, Arlette A. Santiago, Tania E. Soto, Salomón Ramiro Vásquez-García, Daniella Esperanza Pacheco-Catalán, and Joel Vargas. 2023. "Synthesis, Characterization, and Proton Conductivity of Muconic Acid-Based Polyamides Bearing Sulfonated Moieties" Polymers 15, no. 23: 4499. https://doi.org/10.3390/polym15234499

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop