Next Article in Journal
Effect of Fibers on High-Temperature Mechanical Behavior and Microstructure of Reactive Powder Concrete
Previous Article in Journal
Properties of Cement-Based Composites Modified with Magnetite Nanoparticles: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co-Polymers based on Poly(1,4-butylene 2,5-furandicarboxylate) and Poly(propylene oxide) with Tuneable Thermal Properties: Synthesis and Characterization

by
Marina Matos
1,
Andreia F. Sousa
1,*,
Patrícia V. Mendonça
2 and
Armando J. D. Silvestre
1
1
CICECO and Department of Chemistry, University of Aveiro, 3810-193 Aveiro, Portugal
2
CEMMPRE, Department of Chemical Engineering, University of Coimbra, Rua Sílvio Lima, Pólo II, 3030-790 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Materials 2019, 12(2), 328; https://doi.org/10.3390/ma12020328
Submission received: 11 December 2018 / Revised: 15 January 2019 / Accepted: 15 January 2019 / Published: 21 January 2019

Abstract

:
Poly(ether ester)s (PEEs) represent a promising class of segmented co-polymers, nevertheless the synthesis of PEEs based on renewable 2,5-furandicarboxylic acid (FDCA) is still scarce. In this context, a series of poly(1,4-butylene 2,5-furandicarboxylate)-co-poly(poly(propylene oxide) 2,5-furandicarboxylate) co-polyesters with different composition of stiff poly(1,4-butylene 2,5-furandicarboxylate) (PBF) and soft poly(poly(propylene oxide) 2,5-furandicarboxylate) (PPOF) moieties were synthesized, via a two-step bulk polytransesterification reaction. The molar ratio of PBF/PPOF incorporated was varied (10 to 50 mol%) in order to prepare several novel materials with tuned properties. The materials were characterised in detail through several techniques, namely ATR FTIR, 1H and 13C NMR, TGA, DSC, DMTA and XRD. Their hydrolytic and enzymatic degradation evaluation was also assessed. These new co-polymers showed either a semi-crystalline nature when higher PBF/PPOF ratios were used, and for approximately equal amounts of PBF and PPOF an amorphous co-polyester was obtained instead.

1. Introduction

The massive consumption of fossil-based polymers used on a variety of commodity objects of daily life has prompted, in the last decades, the development of renewable-based alternatives with emphasis on their sustainability. Among the renewable-based polymers, polyesters derived from 2,5-furandicarboxylic acid (FDCA) are some of the most promising. FDCA is a renewable-based aromatic building block monomer that has been widely explored as precursor of several homopolyesters, with very similar properties to those of their non-renewable counterparts, namely terephthalic acid-derivatives (TPA) [1,2].
Some examples of polyesters synthesized from FDCA are poly(ethylene 2,5-furandicarboxylate) (PEF) [3] and poly(1,4-butylene 2,5-furandicarboxylate) (PBF), among others, with similar properties to those of their TPA counterparts [4,5,6,7,8,9,10,11] or even enhanced properties such as for example barrier properties or degradability under enzymatic conditions [12,13]. Furthermore, several other diols besides linear ones have been explored to prepare FDCA-based homopolymers, namely branched diols such as 1,2-propanediol, 2,3-butanediol, 2-methyl-1,3-propanediol and 2,2-dimethyl-1,3-propanediol [14,15,16,17,18,19,20]. The ensuing materials presented even higher Tgs than those of their corresponding homopolyesters synthesized with linear diols with the same number of carbon atoms.
In addition, a demand for new polyesters with specific properties emerged due to the necessity to fulfil specific application gaps, namely in the fields of biomaterials or elastomeric compounds. In this context, some studies [11,21,22,23,24] have been focused on the synthesis of FDCA-based poly(ester-ether)s (PEE’s) co-polymers composed of polyether soft moieties, for example, poly(butylene glycol) (PBG) [24] and/or poly(ethylene glycol) (PEG) [22,23] to replace their fossil-based counterparts, such as poly(butylene 1,4-terephthalate)-co-poly(poly(butylene glycol) 1,4-terephtalate) (PBT-co-PBGT) or poly(butylene 1,4-terephthalate)-co-poly(poly(ethylene glycol) 1,4-terephtalate) (PBT-co-PEGT). Commercial PEEs based on TPA, find important applications among the biomedical field [25,26]. Indeed, PBT-co-PEGT co-polymers are commercial PEEs (under trade mark of PolyActive®) [27], and are widely used for drug delivery systems, presenting high thermal stability, as well as enhanced flexibility when compared to PET. This class of polymers show great potential, especially if prepared from FDCA, nonetheless, the literature in furanic PEEs is still scarce [11,21,22,23,24]. Zhou et al. [21] presented the first study, reporting the synthesis of poly(1,4-butylene 2,5-furandicarboxylate)-co-poly(poly(butylene glycol) 2,5-furandicarboxylate) (PBF-co-PBGF) co-polyesters, with enhanced thermal (maximum decomposition temperatures varying from 363 to 378 °C) and mechanical properties (elongation at break between 381 to 832%). Sousa et al. [11] also reported the synthesis of new poly(ester-ether)s co-polymers from FDCA and PEG with different molecular weights (Mn of ~ 200, 400 and 2000), and isosorbide, possessing high thermal stability. More recently, a series of poly(1,4-butylene 2,5-furandicarboxylate)-co-poly(poly(ethylene glycol) 2,5-furandicarboxylate) (PBF-co-PEGF) co-polyesters were also reported, displaying high thermal stability (up to 380 °C) and low Tgs (ranging from −43.1 to −35.4 °C), enabling their manufacturing process [22,23]. Hydrolytic degradability of PBF-b-PEGF was also studied, showing that they can be easily hydrolyzed under alkaline conditions (in phosphate buffered saline (PBS) solution at pH = 12) [23]. Further, Chi et al. [24] also synthesized several PEE’s from FDCA, neopentyl glycol and poly(butylene glycol), with enhanced flexibility (elongation at break values from 38 to 1281%).
However, to our knowledge, the use of poly(propylene oxide) (PPO) as co-monomer for the synthesis of PPE co-polymers has not been reported in the literature before. In this context, co-polymerization of FDCA, 1,4-butanediol (BD) and PPO, could be an elegant way to obtain materials with low Tg, facilitating its processability and, at same time, maintaining the thermal stability, thus enlarging the role of potential applications. Precisely, in this study, a series of poly(1,4-butylene 2,5-furandicarboxylate)-co-poly(poly(propylene oxide) 2,5-furandicarboxylate) (PBF-co-PPOF) co-polyesters were synthesized for the first time by a typical two-step bulk polytransesterification procedure. The ensuing co-polyesters were extensively characterized by SEC, ATR FTIR, 1H and 13C NMR, TGA, DSC, and DMTA analyses. Importantly, PBF-co-PPOF bearing 20 mol% of PPOF moieties was also submitted to a hydrolytic and enzymatic degradability evaluation.

2. Materials and Methods

2.1. Materials

1,4-Butanediol (BD, 99%), poly(propylene oxide) (PPO, average Mn ~1000), deuterated chloroform (CDCl3-d, 99 atom% D), titanium(IV) tert-butoxide (Ti(OBu)4, 97%), poly(methyl methacrylate) standards with molecular weights between 4250 and 273,000 sodium phosphate dibasic (≥99%), sodium phosphate monobasic (≥99%) and Porcine pancreas lipase (Type II, 100-500 units/mg protein) were purchased from Sigma-Aldrich Chemicals Co. 2,5-Furandicarboxylic acid (FDCA, >98%) was purchased from TCI Europe NV. Concentrated hydrochloric acid (37%) was purchased from Panreac; and methanol and chloroform (pro-analysis and HPLC grade) were purchased from Fisher Scientific. All chemicals were used as received, without further purification.

2.2. Synthesis of Dimethyl 2,5-Furandicarboxylate Monomer

The synthesis of DMFDC followed a previously reported procedure [3]. Typically, DMFDC was prepared by reacting FDCA with an excess of methanol, under acidic conditions (HCl), at 80 °C for 15 h. The final product was isolated in 71% yield as a white powder. FTIR (ν/cm−1): 3168 (=C–H); 2965 (C–H); 1706 (C=O); 1578, 1522 (C=C); 1288 (C–O); 1024 (furan ring breathing); 969, 825, 757 (2,5-dibustituted furan ring). 1H NMR (300 MHz, CDCl3, δ/ ppm): 7.2 (s, H3/H4 furan ring); 3.9 (s, 2,5-COOCH3). 13C NMR (75 MHz, CDCl3, δ/ppm): 158 (2,5-C=O); 147 (C2/C5 furan ring); 118 (C3/C4 furan ring); 52 (2,5-COOCH3).

2.3. Syntheses of Poly(1,4-Butylene 2,5-Furandicarboxylate)-Co-Poly(Poly(Propylene Oxide) Co-Polymers and Poly(1,4-Butylene 2,5-Furandicarboxylate)

The polyesters were prepared by a two-step bulk polytransesterification approach following an adapted procedure reported elsewhere [11]. Reactions were carried out by mixing DMFDC (ca. 0.56 g, 3.04 mmol) and a slight excess amount of BD and PPO (ca. 3.22 mmol) (BD/PPO molar ratios of 100/0, 90/10, 80/20, 50/50, Table 1), in the presence Ti(OBu)4 as catalyst (1 wt% relative to the total mass of monomers). In the first step, the mixture was heated progressively from 100 up to 190 °C, for 5 h, under a nitrogen atmosphere and with constant stirring. In the second step, the reaction proceeded under vacuum (10−3 bar) and the temperature was slowly raised to 200 °C for 1h, and finally kept at 210 °C, for 2 h. The ensuing solid products were purified by dissolving in chloroform (~20 mL), and then pouring in an excess of cold methanol, filtered, dried and weighted. In the case of PBF-co-PPOF 50/50 co-polyester (BD/PPO = 50/50), viscous liquid at room temperature, a liquid-liquid extraction procedure using chloroform (~20 mL) was used instead.
Hereafter, the co-polyesters will be referred to as PBF-co-PPOF 90/10, 80/20 and 50/50, according to the BD/PPO molar ratio used as feed. Table 1 presents the molar amounts of each monomer used as well as the weight average molecular weights (Mw) and dispersity (Đ = Mw/Mn) of the polymers.

2.4. Characterization Methods

Size-exclusion chromatography (SEC) analyses of co-polyesters were performed on a Viscotek (Viscotek TDAmax) (Malvern, Gondomar, Portugal) equipped with a differential viscometer (DV) and right-angle laser-light scattering (RALLS, Viscotek) and refractive index (RI) detectors. The column set consisted of a PLgel 5 μm guard column followed by two columns, namely Viscotek T5000 and T4000 column, respectively. A dual piston pump was set with a flow rate of 1 mL min−1. The eluent (DMF with 0.03% LiBr) was previously filtered through a 0.2 μm filter. The system was also equipped with an on-line degasser. The analyses were performed at 60 °C using an Elder CH-150 heater. Before injection, the samples were filtered through a PTFE membrane with 0.2 μm pore. The system was calibrated with narrow poly(methyl methacrylate) standards and the molecular weight and dispersity of the polymers were determined by conventional calibration.
Attenuated total reflectance Fourier transform infrared (ATR FTIR) spectra were obtained using a PARAGON 1000 FTIR spectrometer equipped with a single-horizontal Golden Gate ATR cell (Perkin-Elmer, MA, United States). The spectra were recorded after 128 scans, at a resolution of 4 cm−1, within the range of 500 to 4000 cm−1.
1H, 13C nuclear magnetic resonance (NMR) spectra were recorded in CDCl3 using a Bruker AMX 300 spectrometer (Bruker, Madrid, Spain), operating at 300 or 75 MHz, respectively. All chemical shifts (δ) are expressed as parts per million (ppm), downfield from tetramethylsilane (used as the internal standard). Further, the calculation of the real incorporation of BD/PPO ratio (PBF/PPOFreal) the integration areas of OCH2 proton resonance of F-BD diad (δ ≈ 4.40 ppm) and of F-PPO diad (δ ≈ 3.93 ppm) were used, according to the equation: [ A OCH 2 ; F BD / ( A OC H 2 ; F BD + A OC H 2 ; F PPO ) ] / [ A OC H 2 ; F PPO / ( A OC H 2 ; F DB + A OC H 2 ; F PPO ) ] .
The Average PBF sequence length was also calculated by the equation: L n , B F = 1 / [ ( 1 ( P B F r e a l / 100 ) ) ] [28].
Thermogravimetric analyses (TGA) were carried out with a Setaram SETSYS analyzer (Setaram, Caluire, France) equipped with an alumina plate. Thermograms were recorded under a nitrogen flow of 20 mL min−1 and heated at a constant rate of 10 °C min−1 from room temperature up to 800 °C. Thermal decomposition temperatures were taken at 5% weight loss step and at maximum decomposition temperatures from the heated samples (Td,5% and Td,max, respectively).
Differential scanning calorimetry (DSC) thermograms were obtained with a Hitachi DSC7000X calorimeter (Hitachi, Paris, France) equipped with a liquid nitrogen cooling system, using aluminum DSC pans. Scans were carried out under nitrogen with a heating rate of 5 °C min−1 in the temperature range from −90 to 200 °C. Two heating/cooling cycles were repeated. Glass transition temperature (Tg) was determined using the midpoint approach (second heating trace); and cold crystallization (Tcc) and melting (Tm) temperatures were determined as the maximum of the exothermic crystallization peak and the minimum of the melting endothermic peak during the second heating scan, respectively.
Dynamic mechanical thermal analyses (DMTA) were performed using a material pocket accessory with a Tritec 2000 DMA (Triton, WA, United States), operating in the single cantilever mode. Tests were performed at 1 and 10 Hz and the temperature was varied from -100 to 200 °C, at 2 °C min−1. Tg was determined as the maximum peak of tan δ.
X-ray diffraction (XRD) measurements were performed using a Philips X’pert MPD diffractometer operating with CuKα radiation (λ = 1.5405980 Å) (Malvern, Gondomar, Portugal) at 40 kV and 50 mA. Samples were scanned in the 2θ range of 5 to 50°, with a step size of 0.04°, and time per step of 50 s.
In vitro hydrolytic and enzymatic degradation tests were carried out using press-molded square-shape samples (ca. 69–113 mg) of the prepared polyesters and placed into closed containers with phosphate buffer saline solution (PBS) (10 mL) or with a PBS solution (10 mL) containing Porcine pancreas lipase (concentration of 0.1 mg mL−1), for each test, respectively. The specimens were taken out of the related solution at regular intervals (each 7 days), rinsed thoroughly with distilled water, dried at room temperature for 4 h and, weighed. To prevent saturation, both solutions were renewed every 7 days. Each measurement was repeated five times. The weight loss percentage was calculated using the expression: Weight   loss   ( % ) = [ ( W 0 W d ) / W 0 ] × 100 , where, Wo and Wd stand for the specimens’ weights prior and after incubation, respectively.

3. Results and Discussion

3.1. PBF-co-PPOF Co-Polyesters Synthesis and Structural Characterization

In this study the newly prepared poly(ester-ether) co-polymers are based on poly(1,4-butylene 2,5-furandicarboxylate) as rigid unit and on a soft segment derived from poly(propylene oxide) (Scheme 1). Interestingly, the poly(ether) selected has a methyl side group which plays an important role on the structure-properties features, as discussed ahead.
PBF-co-PPOF were synthesized via a two-step conventional melt polytransesterification approach (Scheme 1) [3], in the presence of Ti(OBu)4 catalyst and at relatively moderate temperatures, not exceeding 210 °C, to avoid undesirable side reactions involving the furan moiety (e.g., decarboxylation reactions which are commonly associated to color problem issues) [1]. The resulting polymers were isolated as powders (PBF, PBF-co-PPOF-90/10 and 80/20) or a viscous liquid (PBF-co-PPOF-50/50), in relatively good yields ranging from 65 to 71% (Table 1) in similarity to other FDCA-based polyesters [22]. Furthermore, these co-polyesters showed weight-average molecular weight (Mw) values between 36,700–48,500, and dispersity (Đ) around 2.
Typical ATR FTIR spectra of all PBF-co-PPOF co-polyesters and related PBF homopolyester (Figure 1) displayed two weak bands near 3150 and 3115 cm−1 attributed to the ν C-H bond of the furanic ring. In addition, near 2968, 2930, 2893 and 2868 cm−1 there are four weak bands attributed to the anti-symmetrical and symmetrical stretching modes (ν C–H asym and ν C–H sym, respectively) of the C–H bond of methylene and methyl groups related to the BD and PPO moieties, respectively. Additionally, both PBF and co-polyesters spectra exhibited a very intense band near 1725 cm−1, arising from the C–O stretching vibration, typical of ester groups. Two bands at 1506 and 1573 cm−1, arising from the C–C bond of the furan ring, and C–O–C stretching vibrations appeared at around 1271 cm−1 and the typical vibration modes of 2,5-disubstituted furans were observed at 966, 822, and 769 cm−1 in the case of PBF and PBF-co-PPOF materials. The presence of the abovementioned bands confirmed the success of the polymerization reactions.
The chemical structure characterization of all PBF-co-PPOF co-polyesters and PBF homopolyester was also studied by 1H (Figure 2 and Table 2), 13C NMR (Table 4 and Figure S1) and 2D analysis (Figure S2). The main 1H NMR resonances and respective assignments of all polymers studied are summarized in Table 2 and Figure 2 displays the 1H NMR spectrum of PBF-co-PPOF-90/10.
The 1H NMR spectra of all polymers (Figure 2 and Table 2) displayed the typical resonances attributed to the F-BD diad at approximately δ 7.21, 4.40 and 1.91 attributed to the H3 and H4 protons of the furan ring, and to the C(O)OCH2CH2 and C(O)OCH2CH2 protons of the BD moiety, respectively. In the co-polymers spectra, the corresponding resonances associated to the PPO-F diads were also detected: δ 5.26, 3.67 and 1.34 ppm, arising from the C(O)OCH, C(O)OCH2 and C(O)OCHCH3 protons, in the neighboring of the furan ring, respectively. Moreover, the protons related to the PPO-PPO units were also identified at δ 3.54, 3.40, and 1.15 ppm, related to OCH2 (ether linkage), OCH and OCHCH3, respectively.
Furthermore, the 1H NMR spectra data was used to access the real molar percentage of PBF and PPOF moieties in the co-polyesters backbone, due to the important impact this ratio has on the ensuing co-polyesters properties. The PBF/PPOF real incorporation was determined using the integration areas of C(O)OCH2 proton resonances (in the neighboring of furan ring) of the F-BD (δ at 4.40 ppm) and F-PPO (δ at 3.93 ppm) diads, respectively, and the main results are presented in Table 3.
From Table 3, it is possible to observe that despite the BD and PPO feed ratio, there was a tendency for a higher incorporation of PPOF into co-polyesters chains, most probably associated with BD lost during the polytransesterification step due to the high BD volatility. However, this trend is almost negligible in the case of the PBF-co-PPOF-90/10 and 80/20 co-polyesters.
The number average sequence length of BF unit (Ln,BF) was also assessed assuming that PBF-co-PPOF co-polyesters are random co-polyesters. It was found that Ln,BF increased with the BF content increasing, according with the theoretically expected values [11,22,29,30]. Importantly, the co-polyester with the highest amount of PBF (PBF-co-PPOF-90/10) had a Ln,BF of 6.5. This is an important structural feature that is in accordance with a crystalline domain dominated by PBF segments and corresponding melting behavior, as discussed above.
In terms of 13C NMR analysis (Table 4 and Figure S1), the observed resonances were in agreement with their expected chemical structure and corroborated the above 1H NMR results, as well as the ATR FTIR data.

3.2. X-Ray Diffraction Analysis

The XRD patterns of PBF-co-PPOF-90/10 and 80/20 co-polymers (Figure 3) indicated that they were semi-crystalline polymers, displaying both an amorphous halo around 2θ ~ 21° and diffraction peaks at 2θ ≈ 18, 23 and 25°, quite similar to the PBF pattern although in the former case peaks were more intense than for the co-polymers probably due to a higher crystallinity. In fact, the PBF XRD pattern (Figure 3), exhibited intense diffraction peaks at 2θ ~ 10, 18, 23 and 25° [16,31,32,33]. This clearly indicates that the ability of PBF-co-PPOF co-polyesters to crystallize is mainly associated to PBF segments (with Ln,BF equal to 6.5 and 4.2, respectively). Moreover, these results are in perfect agreement with the below DSC and DMTA data, and also with other FDCA or TPA-based PPEs reported in the literature [21,22,34,35].
As expected, in the case of the viscous liquid PBF-co-PPOF-50/50 co-polyester, only a halo centered at 2θ ≈ 19 ° was observed, in accordance with an essential amorphous nature.

3.3. Thermal Behaviur

PBF-co-PBDG co-polyesters were extensively characterized in terms of their thermal behavior through DSC, DMTA and TGA analyses (Table 5, and Figures S3–S6).
The semi-crystalline character of the co-polymers having higher PBF/PPOF molar ratios (PBF-co-PPOF-90/10 and 80/20) was confirmed by DSC and DMTA analyses (Figures S3–S5). For example, the DSC trace of PBF-co-PPOF-90/10 co-polyester (Figure S4) displayed glass transition (Tg) and melting temperatures (Tm) at −49.7 and 137.7 °C, respectively. The PBF-co-PPOF-80/20 co-polyester also maintained some ability to crystallize and melt, with a cold transition (Tcc) and Tm of approximately 21.7 and 124.1 °C (Table 5), similar to the one of PBF (Tm of 170.0 °C). Furthermore, using DTMA analysis (Figure S5), Tg values of −32.6, −37.2 and −42.3 °C for PBF-co-PPOF-90/10, 80/20 and 50/50 materials were observed, respectively; and Tm of 147.6 and 124.2 °C for PBF-co-PPOF-90/10 and 80/20 were detected. In the case of PBF-co-PPOF-50/50 co-polymer (having a lower PBF/PPOF molar ratio) the absence of Tm is in accordance with its essential amorphous nature (XRD data).
In summary, the highly desirable properties obtained included, the Tg of the co-polyesters decreased with the increasing content of soft PPOF segments (tailored behavior), although these results were still higher than those of amorphous PPO homopolyester (Tg between −67 to −72 °C [36,37]). These results were expected since the incorporation of more PPO flexible moieties into FDCA-based co-polyesters typically gives rise to a decrease on the co-polymers’ thermal features [21,22]. Importantly, the co-polymers with higher BF content, PBF-co-PPOF-90/10 and 80/20, showed a typical segmented polymers behavior [22], with a Tm very close to that of stiff PBF and a Tg below room temperature closer to that of soft PPO, revealing that PBF units were the main responsible for the crystalline behavior of the ensuing materials, whereas the PPOF segments were associated with the amorphous domain. Hence, for PBF-co-PPOF-90/10 the range of working temperatures was quite enlarged: within −39 and 139 °C.
In general the TGA thermograms (Table 5 and Figure S6) of the co-polyesters (carried out under nitrogen atmosphere) exhibited one major characteristic event at the maximum decomposition temperatures (Td,max) of 340−365 °C. Also, the newly prepared co-polymers showed to be thermally stable up to Td,5% ≈ 308 °C.
As shown in Table 5, the co-polymers had both Td,5% and Td,max results lower than those observed to PBF. These less favorable thermal results could be associated with the presence of the appending methyl group, as already reported for other polyesters also having side groups, such as poly(2,3-butylene 2,5-furandicarboxylate) compared to PBF [17]. Nevertheless, all PBF-co-PPOF co-polyesters have higher Td,max than PPO.

3.4. Hydrolytic and Enzymatic Degradation Tests

As mentioned above, in addition to specific thermal properties, it was also important to understand if the newly prepared materials degraded under hydrolytic and enzymatic conditions for a relatively short period of time. The evaluation of the hydrolytic and enzymatic degradation behavior was performed in terms of weight loss percentage versus time (Figure 4) for PBF-co-PPOF-80/20 co-polyester, with real incorporation of PPOF moieties around 24 mol%. However, the polymer’s weight loss under hydrolytic conditions was almost negligible for a relatively short period of 12 weeks, and a little bit higher under enzymatic conditions (2.3%). This result was in the same line with those reported for PBF-co-PEGF co-polyesters incorporating similar amount of PBF moieties [23], and also with studies on PBT-co-PEGT co-polyesters in similar hydrolytic conditions as the ones used in this study (pH ~ 7 and 37 °C) [38].
The incubation with a Porcine pancreas lipase in PBS solution slightly increased the weight loss of the co-polyester, but it was still very low. Through hydrolytic and enzymatic degradation only 1.5 and 2.3% weight loss were achieved, respectively. Nevertheless, the data obtained are good preliminary results, and showed some evidence of degradation of PBF-co-PPOF (also checked by FTIR) associated with the hydrolysis of the ester groups, but further studies using other enzymes such as for example cutinases, could enhance the results obtained. Some positive results were previously obtained with cutinase from Humicola insolens or Thermobifida cellulosilytica to hydrolyze PET and PEF [12,13]. Further, incorporation of PPO in the co-polymers, is also, expected to enhance degradability.

5. Conclusions

In conclusion, a new class of FDCA-based poly(ester-ether) co-polymers has been accomplished, incorporating both stiff and soft moieties into their backbone. The ensuing co-polyesters have shown high thermal stability (Td,max between 340 to 365 °C) and Tg at sub-ambient temperatures, namely from −42.3 to −32.6 °C. Moreover, the semi-crystalline character was only observed for co-polyesters with higher BF content, revealing that PBF units were mainly responsible for the crystalline behavior of the ensuing materials, whereas the PPOF were associated with the amorphous domain. Furthermore, PBF-co-PPOF-80/20 co-polyester had shown a week hydrolysable behavior, presenting a maximum percentage weight loss of 2.3 %, after 12 weeks.
Finally, due to their high thermal stability, as well as the presence of both stiff and soft moieties in the co-polymer chains, these materials could find interesting industrial applications, namely as thermoplastic polymers.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/12/2/328/s1, Figure S1: 13C NMR spectra in CHCl3-d of PBF-co-PPOF-90/10 copolymer, Figure S2: 2D NMR spectrum in CHCl3-d of PBF-co-PPOF-90/10 copolymer, Figure S3: DSC curve of PBF homopolyester. Figure S4: DSC curves (second scan) of all PBF-co-PPOF copolymers studied, Figure S5: Tan δ of PBF and PBF-co-PPOF (co)polymers, at 1 Hz, Figure S6: TGA (a) and DTGA (b) thermograms of PBF-co-PPOF copolymers and PBF homopolyester.

Author Contributions

Supervision, A.J.D.S. and A.F.S.; conceptualization, A.F.S.; investigation, M.M.; SEC analysis performance, P.V.M.

Acknowledgments

FCT and POPH/FSE are gratefully acknowledged for funding a doctoral grant to MM (PD/BD/52501/2014) and the post-doctoral grant of PVM (SFRH/BPD/117589/2016). The costs resulting from the FCT hiring of AFS was funded by national funds (OE), through FCT – Fundação para a Ciência e a Tecnologia , I.P., in the scope of the framework contract foreseen in the numbers 4, 5 and 6 of the article 23, of the Decree-Law 57/2016, of August 29, changed by Law 57/2017, of July 19. This work was developed within the scope of the project CICECO-Aveiro Institute of Materials, FCT Ref. UID/CTM/50011/2019, financed by national funds through the FCT/MCTES.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. De Jong, E.; Dam, M.A.; Sipos, L.; Gruter, G.-J. Furandicarboxylic acid (FDCA), a versatile building block for a very interesting class of polyesters. In Biobased Monomers, Polymers, and Materials; Smith, P.B., Gross, R.A., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2012; Volume 1105, pp. 1–13. ISBN 0-8412-2767-5. [Google Scholar]
  2. Jia, Z.; Wang, J.; Sun, L.; Zhu, J.; Liu, X. Fully bio-based polyesters derived from 2,5-furandicarboxylic acid (2,5-FDCA) and dodecanedioic acid (DDCA): From semicrystalline thermoplastic to amorphous elastomer. J. Appl. Polym. Sci. 2018, 135, 46076. [Google Scholar] [CrossRef]
  3. Gandini, A.; Silvestre, A.J.D.; Neto, C.P.; Sousa, A.F.; Gomes, M. The furan counterpart of poly(ethylene terephthalate): An alternative material based on renewable resources. J. Polym. Sci. Polym. Chem. 2009, 47, 295–298. [Google Scholar] [CrossRef] [Green Version]
  4. Sousa, A.F.A.F.; Vilela, C.; Fonseca, A.C.; Matos, M.; Freire, C.S.R.; Gruter, G.-J.M.; Coelho, J.F.J.; Silvestre, A.J.D. Biobased polyesters and other polymers from 2,5-furandicarboxylic acid: A tribute to furan excellency. Polym. Chem. 2015, 6, 5961–5983. [Google Scholar] [CrossRef]
  5. Zhang, D.; Dumont, M.J. Advances in polymer precursors and bio-based polymers synthesized from 5-hydroxymethylfurfural. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 1478–1492. [Google Scholar] [CrossRef]
  6. Papageorgiou, G.Z.; Papageorgiou, D.G.; Terzopoulou, Z.; Bikiaris, D.N. Production of bio-based 2,5-furan dicarboxylate polyesters: Recent progress and critical aspects in their synthesis and thermal properties. Eur. Polym. J. 2016, 83, 202–229. [Google Scholar] [CrossRef]
  7. Gomes, M.; Gandini, A.; Silvestre, A.J.D.; Reis, B. Synthesis and characterization of poly(2,5-furan dicarboxylate)s based on a variety of diols. J. Polym. Sci. Polym. Chem. 2011, 49, 3759–3768. [Google Scholar] [CrossRef]
  8. Sousa, A.F.; Matos, M.; Freire, C.S.R.; Silvestre, A.J.D.; Coelho, J.F.J. New copolyesters derived from terephthalic and 2,5-furandicarboxylic acids: A step forward in the development of biobased polyesters. Polymer 2013, 54, 513–519. [Google Scholar] [CrossRef]
  9. Matos, M.; Sousa, A.F.; Silvestre, A.J.D. Improving the thermal properties of poly(2,5-furandicarboxylate)s using cyclohexylene moieties: A comparative study. Macromol. Chem. Phys. 2017, 218, 1600492. [Google Scholar] [CrossRef]
  10. Soares, M.J.; Dannecker, P.-K.; Vilela, C.; Bastos, J.; Meier, M.A.R.; Sousa, A.F. Poly(1,20-eicosanediyl 2,5-furandicarboxylate), a biodegradable polyester from renewable resources. Eur. Polym. J. 2017, 90, 301–311. [Google Scholar] [CrossRef]
  11. Sousa, A.F.; Coelho, J.F.J.; Silvestre, A.J.D. Renewable-based poly((ether)ester)s from 2,5-furandicarboxylic acid. Polymer 2016, 98, 129–135. [Google Scholar] [CrossRef]
  12. Weinberger, S.; Canadell, J.; Quartinello, F.; Yeniad, B.; Arias, A.; Pellis, A.; Guebitz, G. Enzymatic degradation of poly(ethylene 2,5-furanoate) powders and amorphous films. Catalysts 2017, 7, 318. [Google Scholar] [CrossRef]
  13. Weinberger, S.; Haernvall, K.; Scaini, D.; Ghazaryan, G.; Zumstein, M.T.; Sander, M.; Pellis, A.; Guebitz, G.M. Enzymatic surface hydrolysis of poly(ethylene furanoate) thin films of various crystallinities. Green Chem. 2017, 19, 5381–5384. [Google Scholar] [CrossRef]
  14. Haernvall, K.; Zitzenbacher, S.; Amer, H.; Zumstein, M.T.; Sander, M.; McNeill, K.; Yamamoto, M.; Schick, M.B.; Ribitsch, D.; Guebitz, G.M. Polyol structure influences enzymatic hydrolysis of bio-based 2,5-furandicarboxylic acid (FDCA) polyesters. Biotechnol. J. 2017, 12, 1600741. [Google Scholar] [CrossRef] [PubMed]
  15. Gubbels, E.; Jasinska-Walc, L.; Koning, C.E. Synthesis and characterization of novel renewable polyesters based on 2,5-furandicarboxylic acid and 2,3-butanediol. J. Polym. Sci. Polym. Chem. 2013, 51, 890–898. [Google Scholar] [CrossRef]
  16. Thiyagarajan, S.; Vogelzang, W.; Knoop, R.J.; Frissen, A.E.; van Haveren, J.; van Es, D.S. Biobased furandicarboxylic acids (FDCAs): Effects of isomeric substitution on polyester synthesis and properties. Green Chem. 2014, 16, 1957. [Google Scholar] [CrossRef]
  17. Jiang, Y.; Woortman, A.J.J.; van Ekenstein, G.O.R.A.; Loos, K. A biocatalytic approach towards sustainable furanic-aliphatic polyesters. Polym. Chem. 2015, 6, 5198–5211. [Google Scholar] [CrossRef]
  18. Terzopoulou, Z.; Kasmi, N.; Tsanaktsis, V.; Doulakas, N.; Bikiaris, D.N.; Achilias, D.S.; Papageorgiou, G.Z. Synthesis and characterization of bio-based polyesters: Poly(2-methyl-1,3-propylene-2,5-furanoate), Poly(isosorbide-2,5-furanoate), Poly(1,4-cyclohexanedimethylene-2,5-furanoate). Materials 2017, 10, 801. [Google Scholar] [CrossRef]
  19. Tsanaktsis, V.; Terzopoulou, Z.; Exarhopoulos, S.; Bikiaris, D.N.; Achilias, D.S.; Papageorgiou, D.G.; Papageorgiou, G.Z. Sustainable, eco-friendly polyesters synthesized from renewable resources: Preparation and thermal characteristics of poly(dimethyl-propylene furanoate). Polym. Chem. 2015, 6, 8284–8296. [Google Scholar] [CrossRef]
  20. Genovese, L.; Lotti, N.; Siracusa, V.; Munari, A. Poly(neopentyl glycol furanoate): A member of the furan-based polyester family with smart barrier performances for sustainable food packaging applications. Materials 2017, 10, 1028. [Google Scholar] [CrossRef]
  21. Zhou, W.; Zhang, Y.; Xu, Y.; Wang, P.; Gao, L.; Zhang, W.; Ji, J. Synthesis and characterization of bio-based poly(butylene furandicarboxylate)-b-poly(tetramethylene glycol) copolymers. Polym. Degrad. Stab. 2014, 109, 21–26. [Google Scholar] [CrossRef]
  22. Sousa, A.F.; Guigo, N.; Pozycka, M.; Delgado, M.; Soares, J.; Mendonça, P.V.; Coelho, J.F.J.; Sbirrazzuoli, N.; Silvestre, A.J.D. Tailored design of renewable copolymers based on poly(1,4-butylene 2,5-furandicarboxylate) and poly(ethylene glycol) with refined thermal properties. Polym. Chem. 2018, 9, 722–731. [Google Scholar] [CrossRef]
  23. Hu, H.; Zhang, R.; Sousa, A.; Long, Y.Y.; Ying, W.B.; Wang, J.; Zhu, J. Bio-based poly(butylene 2,5-furandicarboxylate)-b-poly(ethylene glycol) copolymers with adjustable degradation rate and mechanical properties: Synthesis and characterization. Eur Polym J 2018, 106, 42–52. [Google Scholar] [CrossRef]
  24. Chi, D.; Liu, F.; Na, H.; Chen, J.; Hao, C.; Zhu, J. Poly(neopentyl glycol 2,5-furandicarboxylate): A Promising Hard Segment for the Development of Bio-based Thermoplastic Poly(ether-ester) Elastomer with High Performance. ACS Sustain. Chem. Eng. 2018, 6, 9893–9902. [Google Scholar] [CrossRef]
  25. Deschamps, A.A.; Claase, M.B.; Sleijster, W.J.; Bruijn, J.D.; Grijpma, D.W.; Feijen, J. Design of segmented poly(ether ester) materials and structures. J. Control. Release 2002, 78, 175–186. [Google Scholar] [CrossRef]
  26. Buitinga, M.; Truckenmüller, R.; Engelse, M.A.; Moroni, L.; Ten Hoopen, H.W.; van Blitterswijk, C.A.; de Koning, E.J.; van Apeldoorn, A.A.; Karperien, M. Microwell Scaffolds for the Extrahepatic Transplantation of Islets of Langerhans. PLoS ONE 2013, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  27. Octoplus PolyActiveTM A Biodegradable Polymer-Based Drug Delivery System. Available online: http://www.octoplus.nl/nl/Contact/News (accessed on 3 December 2018).
  28. Deschamps, A.A.; Grijpma, D.W.; Feijen, J. Poly(ethylene oxide)/poly(butylene terephthalate) segmented block copolymers: The effect of copolymer composition on physical properties and degradation behavior. Polymer 2001, 42, 9335–9345. [Google Scholar] [CrossRef]
  29. Wang, J.; Liu, X.; Jia, Z.; Liu, Y.; Sun, L.; Zhu, J. Synthesis of bio-based poly(ethylene 2,5-furandicarboxylate) copolyesters: Higher glass transition temperature, better transparency, and good barrier properties. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 3298–3307. [Google Scholar] [CrossRef]
  30. Wang, J.; Liu, X.; Zhu, J.; Jiang, Y. Copolyesters based on 2,5-furandicarboxylic acid (FDCA): Effect of 2,2,4,4-tetramethyl-1,3-cyclobutanediol units on their properties. Polymers 2017, 9, 305. [Google Scholar] [CrossRef]
  31. Gandini, A.; Coelho, D.; Gomes, M.; Reis, B.; Silvestre, A. Materials from renewable resources based on furan monomers and furan chemistry: Work in progress. J. Mater. Chem. 2009, 19, 8656. [Google Scholar] [CrossRef]
  32. Jiang, M.; Liu, Q.; Zhang, Q.; Ye, C.; Zhou, G. A series of furan-aromatic polyesters synthesized via direct esterification method based on renewable resources. J. Polym. Sci. Polym. Chem. 2012, 50, 1026–1036. [Google Scholar] [CrossRef]
  33. Ma, J.; Yu, X.; Xu, J.; Pang, Y. Synthesis and crystallinity of poly(butylene 2,5-furandicarboxylate). Polymer 2012, 53, 4145–4151. [Google Scholar] [CrossRef] [Green Version]
  34. Yao, C.; Yang, G. Crystallization, and morphology of poly(trimethylene terephthalate)/poly(ethylene oxide terephthalate) segmented block copolymers. Polymer 2010, 51, 1516–1523. [Google Scholar] [CrossRef]
  35. Szymczyk, A. Structure and properties of new polyester elastomers composed of poly(trimethylene terephthalate) and poly(ethylene oxide). Eur. Polym. J. 2009, 45, 2653–2664. [Google Scholar] [CrossRef]
  36. Allen, G.; Crossley, H.G. Polypropylene oxide IV-Preparation and properties of polyether networks. Polymer 1964, 5, 553–557. [Google Scholar] [CrossRef]
  37. Vesterinen, A.; Lipponen, S.; Rich, J.; Seppälä, J. Effect of block composition on thermal properties and melt viscosity of poly[2-(dimethylamino)ethyl methacrylate], poly(ethylene oxide) and poly(propylene oxide) block co-polymers. Express Polym. Lett. 2011, 5, 754–765. [Google Scholar] [CrossRef]
  38. Chao, G.T.; Fan, L.Y.; Jia, W.J.; Qian, Z.Y.; Gu, Y.C.; Liu, C.B.; Ni, X.P.; Li, J.; Deng, H.X.; Gong, C.Y.; et al. Synthesis, characterization and hydrolytic degradation of degradable poly(butylene terephthalate)/poly(ethylene glycol) (PBT/PEG) copolymers. J. Mater. Sci. Mater. Med. 2007, 18, 449–455. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of PBF-co-PPOF co-polymers.
Scheme 1. Synthesis of PBF-co-PPOF co-polymers.
Materials 12 00328 sch001
Figure 1. ATR FTIR spectra of PBF-co-PPOF co-polymers.
Figure 1. ATR FTIR spectra of PBF-co-PPOF co-polymers.
Materials 12 00328 g001
Figure 2. 1H NMR spectrum of PBF-co-PPOF-90/10 co-polymer in CHCl3-d.
Figure 2. 1H NMR spectrum of PBF-co-PPOF-90/10 co-polymer in CHCl3-d.
Materials 12 00328 g002
Figure 3. XRD patterns of PBF and all PBF-co-PPOF materials.
Figure 3. XRD patterns of PBF and all PBF-co-PPOF materials.
Materials 12 00328 g003
Figure 4. Weight loss percentage of PBF-co-PPOF-80/20 co-polyester along 12 weeks.
Figure 4. Weight loss percentage of PBF-co-PPOF-80/20 co-polyester along 12 weeks.
Materials 12 00328 g004
Table 1. Molecular characteristics of PBF and PBF-co-PPOF co-polymers.
Table 1. Molecular characteristics of PBF and PBF-co-PPOF co-polymers.
(co)polymerBD/PPOfeed 1 (mol%)Yield (%) 2Mw3Đ4
PBF100/071.0-- 5
PBF-co-PPOF-90/1090/1064.836 7002.2
PBF-co-PPOF-80/2080/2068.141 6002.2
PBF-co-PPOF-50/5050/5071.448 5002.1
1 Molar feed percentage of BD and PPO units. 2 Isolation yields of purified polyesters in methanol. 3 Weight-average molecular weight. 4 Dispersity. 5 Not determinate due to the insolubility of PBF in DMF.
Table 2. Main 1H NMR resonances of PBF-co-PPOF co-polyesters, and PBF and PPO homopolyesters. The a, b, c, d, e, f, g and a′, b′, c′, d′, e′ attributions are explained in the scheme of Figure 2.
Table 2. Main 1H NMR resonances of PBF-co-PPOF co-polyesters, and PBF and PPO homopolyesters. The a, b, c, d, e, f, g and a′, b′, c′, d′, e′ attributions are explained in the scheme of Figure 2.
δ/ppmAssignmentDiadsIntegration Area
PBFPBF-co-PPOFPPO
90/1080/2050/50
7.21H3 and H4; CHF-BD; FDCA-PPO1.001.001.001.00
5.26d, d′; CHCH3F-PPO0.140.370.92
4.40a, a′; C(O)OCH2CH2F-BD2.011.661.280.46
3.67c, c′; C(O)OCH2F-PPO0.300.841.84
3.54f; OCH2PPO-PPO1.915.1417.892.01
3.40g; OCHPPO-PPO1.032.779.461.00
1.91b, b′; C(O)OCH2CH2F-BD2.011.661.410.48
1.34e, e′; CHCH3F-PPO 0.441.303.04
1.15h; OCHCH3PPO-PPO3.108.6028.503.02
Table 3. Comparison between the initial and the real molar ratio percentages of PBF and PPOF in the polyesters’ backbone.
Table 3. Comparison between the initial and the real molar ratio percentages of PBF and PPOF in the polyesters’ backbone.
(co)polymerPBF/PPOFfeed 1PBF/PPOFreal 2Ln,BF3
PBF100/0100/0
PBF-co-PPOF-90/1090/1085/156.5
PBF-co-PPOF-80/2080/2076/244.2
PBF-co-PPPOF-50/5050/5020/801.3
1 Molar feed ratio percentage of BD and PPO units. 2 Molar real ratio percentage of BD and PPO units. 3 Average PBF sequence length.
Table 4. Main 13C NMR resonances of all PBF-co-PPOF co-polyesters, and PBF and PPO homopolyester. The a, b, c, d, e, f, g and a′, b′, c′, d′, e′ attributions are explained in the scheme of Figure 2.
Table 4. Main 13C NMR resonances of all PBF-co-PPOF co-polyesters, and PBF and PPO homopolyester. The a, b, c, d, e, f, g and a′, b′, c′, d′, e′ attributions are explained in the scheme of Figure 2.
AssignmentDiadsIntegration Area
PBFPBF-co-PPOFPPO
90/1080/2050/50
2-CO and 5-CO; C(O)OF-BD; F-PPO158.0158.0158.0158.0-
C2 and C5; C-C(O)OF-BD; F-PPO146.8146.8146.8147.0-
C3 and C4; C-HF-BD; F-PPO118.5118.5118.5118.2-
f; OCH2PPO-PPO-75.475.375.375.3
g; OCHPPO-PPO-73.473.473.474.4
c, c′; C(O)OCH2F-PPO-72.972.972.9-
d, d′; CHCH3F-PPO-71.771.771.3-
a, a′; C(O)OCH2CH2F-BD64.964.964.964.8-
b, b′; C(O)OCH2CH2F-BD25.425.425.325.0-
h ; OCHCH3PPO-PPO -17.317.317.317.3
e, e′; CHCH3F-PPO-16.916.816.8-
Table 5. Decomposition at 5% weight loss (Td,5%), maximum decomposition (Td,max), glass transition (Tg) and melting temperature (Tm) of PBF, PPO and PBF-co-PPOF co-polyesters.
Table 5. Decomposition at 5% weight loss (Td,5%), maximum decomposition (Td,max), glass transition (Tg) and melting temperature (Tm) of PBF, PPO and PBF-co-PPOF co-polyesters.
co(polymer)DSC 1DMTA 2TGA 3
Tm/°CTg/°CTm/°CTd,5%/°CTd,max/°C
PBF170.075.6166.9348.7380.5
PBF-co-PPOF-90/10137.7−32.6147.6284.7347.2
PBF-co-PPOF-80/20124.1−37.2124.2288.9340.3
PBF-co-PPOF-50/50-−42.3- 3308.1365.2
PPO −67.0–72.0 5-283.0323.4
1 Determined by DSC analysis. 2 Determined by DMTA analysis. 3 Determined by TGA analysis. 4 Values obtained from references [36,37].

Share and Cite

MDPI and ACS Style

Matos, M.; Sousa, A.F.; Mendonça, P.V.; Silvestre, A.J.D. Co-Polymers based on Poly(1,4-butylene 2,5-furandicarboxylate) and Poly(propylene oxide) with Tuneable Thermal Properties: Synthesis and Characterization. Materials 2019, 12, 328. https://doi.org/10.3390/ma12020328

AMA Style

Matos M, Sousa AF, Mendonça PV, Silvestre AJD. Co-Polymers based on Poly(1,4-butylene 2,5-furandicarboxylate) and Poly(propylene oxide) with Tuneable Thermal Properties: Synthesis and Characterization. Materials. 2019; 12(2):328. https://doi.org/10.3390/ma12020328

Chicago/Turabian Style

Matos, Marina, Andreia F. Sousa, Patrícia V. Mendonça, and Armando J. D. Silvestre. 2019. "Co-Polymers based on Poly(1,4-butylene 2,5-furandicarboxylate) and Poly(propylene oxide) with Tuneable Thermal Properties: Synthesis and Characterization" Materials 12, no. 2: 328. https://doi.org/10.3390/ma12020328

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