Synthesis, Thermal Properties, and Rheological Characteristics of

16 hours ago - Currently, there is an intensive development of bio-based aromatic building blocks to replace fossil-based terephthalates used for ...
0 downloads 0 Views 2MB Size
This is an open access article published under a Creative Commons Attribution (CC-BY) License, which permits unrestricted use, distribution and reproduction in any medium, provided the author and source are cited.

Article Cite This: ACS Omega XXXX, XXX, XXX−XXX

http://pubs.acs.org/journal/acsodf

Synthesis, Thermal Properties, and Rheological Characteristics of Indole-Based Aromatic Polyesters Carlos R. Arza and Baozhong Zhang* Centre of Analysis and Synthesis, Lund University, P.O. Box 124, SE-22100 Lund, Sweden

Downloaded via 185.89.100.37 on September 5, 2019 at 20:01:02 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Currently, there is an intensive development of bio-based aromatic building blocks to replace fossil-based terephthalates used for poly(ethylene terephthalate) production. Indole is a ubiquitous aromatic unit in nature, which has great potential as a bio-based feedstock for polymers or plastics. In this study, we describe the synthesis and characterization of new indole-based dicarboxylate monomers with only aromatic ester bonds, which can improve the thermal stability and glasstransition temperature (Tg) of the resulting polyesters. The new dicarboxylate monomers were polymerized with five aliphatic diols to yield 10 new polyesters with tunable chemical structures and physical properties. Particularly, the Tg values of the obtained polyesters can be as high as 113 °C, as indicated by differential scanning calorimetry and dynamic mechanical analysis. The polyesters showed decent thermal stability and distinct flow transitions as revealed by thermogravimetric analysis and rheology measurements.



INTRODUCTION Today, the production of plastics (polymers) is facing major challenges, which include, for example, the deployment of fossil resources, the rise of petrochemical prices, and the environmental issues associated with the polymer production. Therefore, there is a strong motivation from the society and industry to switch to sustainable bio-based feedstocks for the production of plastics.1 To date, the most successful development of bio-based plastics is Coca-Cola’s partially bio-based poly(ethylene terephthalate) (PET) (PlantBottle),2 which constitutes approximately 75% of all of the bioplastics on the market. Unfortunately, this bio-PET only contains ∼30% mass originated from bio-based production (i.e., biobased ethylene glycol), while the remaining ∼70% mass was still based on the use of fossil-based terephthalic acid (TPA) or dimethyl terephthalate (DMT). The production of TPA or DMT from renewable starting materials has been an active research area, which has achieved lab-scale success.3−12 Other benzenoid aromatic monomers (e.g., dicarboxylates or monohydroxycarboxylates) have also been synthesized and used for the production of aromatic polyesters as a potential substitution for PET, and their raw materials include, for example, lignin-derived molecules,13−15 vanillic acid,16−19 resorcinol,20 and cinnamic acid.21−24 In the meantime, alternative nonbenzenoid aromatic building blocks have also received growing attention, among which furan derivatives have been intensively explored. Furan structures exist widely in nature and can be produced via different biosynthetic pathways. Particularly, furandicarboxylic acid (FDCA) has been a popular research topic recently, because its derived polyester (e.g., poly(ethylene © XXXX American Chemical Society

furanoate) or PEF) has comparable mechanical and even better barrier properties compared with PET.25−33 Other furan-based dicarboxylates and polyesters have also been reported in the literature.28 Unfortunately, the thermal instability of furan derivatives (e.g., FDCA, or precursors like 5-hydroxymethylfurfural, or 5-hydroxymethyl-2-furoic acid) has caused undesirable degradation and coloration during the production and processing at high temperatures (>200 °C), which hinders the further development of furan-based polyesters.34 Recently, a novel procedure to prepare bottlegrade PEF was reported, using a three-step synthesis including polycondensation, oligomerization, and ring-opening polymerization.35 However, this procedure is complicated and its feasibility for the plastic industry remains to be explored. Considering these challenges, it is time to consider other bio-based aromatics as “Plan B” for polyester production. Recently, indole has attracted our attention because it is a large aromatic unit that exists widely in nature and urban wastes. Indole is also a major byproduct in coal mining, petroleum, tobacco, and livestock farming industries36 and is widely used in pharmaceuticals, agrochemicals, and material sciences.37−40 Several bioproduction methods for indole are available, including thermocatalytic conversion and ammonization of biomass-derived furfural and furan, microalgae pyrolysis, yeast and bacteria fermentation, reduction of indigo that is derived from plants and microbial production, and direct conversion from bio-based ethylene glycol and aniline.30−52 Received: June 18, 2019 Accepted: August 7, 2019

A

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

bonds in the new series of polyesters clearly enhanced the thermal stability.

Surprisingly, little attention has been paid to the development of indole-based polyesters. In the literature, the use of indole structures has been limited to conducting polymers and microporous materials,53−56 polymers with indole pendant groups, polyethersulfones, and poly(N-arylene diindolylmethane)s.57−63 There was only one report in the 1980s describing bis-indole-containing polyesters without the characterization of properties.64 Recently, a new series of fully biobased polyesters were synthesized and reported by our group using a new indole-based dicarboxylate monomer and five different aliphatic diols.65 The obtained indole-based polyesters were all amorphous with Tg values of up to 99 °C. This clearly demonstrated the feasibility of making indole-based polyesters and its potential to replace fossil-based terephthalate. A remaining question to be addressed is that the reported series of indole-based polyesters contain 50% aliphatic ester bonds (i.e., esters derived from aliphatic carboxylic acids, structure a, Figure 1), which are not present



RESULTS AND DISCUSSION Polymer Synthesis. Methyl indole-3-carboxylate (1) was employed as the starting material for monomer synthesis, which can be conveniently synthesized from sustainable raw materials, e.g., by a base-mediated carboxylation of indole, followed by transesterification with methanol.66 A possible synthetic route for 1 from biosourced raw materials is depicted in Figure S1, Supporting Information. Dicarboxylate monomers (3a,b) were synthesized by a simple SN2 reaction of commercially available methyl indole-3-carboxylate 1 and two different 1,ω-dibromoalkanes (2a,b), respectively (Scheme 1). Purification of the crude products 3a,b by straightforward recrystallization from toluene afforded white crystals of monomers with >80% yields. Afterward, monomers 3a,b were polymerized with five potentially bio-based aliphatic diols (4a−e), according to a conventional bulk polycondensation protocol using dibutyltin(IV) oxide (DBTO) catalyst (1 mol %).65,67,68 The polymerization was performed in two stages. In the first stage (transesterification), an excess of diol (3 equiv of OH groups for 4c−e and 4 equiv of OH groups for 4a,b) was reacted with dicarboxylates (3a,b) at 180 °C under a slow nitrogen flow to remove the condensed methanol. Once the transesterification was completed (ca 3 h, according to 1H NMR analysis), a second stage (polycondensation) was carried out at 220 °C with a stronger N2 gas flow to completely remove the condensed diols. In the literature, the removal of the condensed diols was usually achieved under high vacuum conditions.69 However, when the vacuum condition was applied for our polymerizations, the obtained polymers always showed an intense red color, while our new protocol under a N2 gas flow provided white or colorless polymers (see examples in the Supporting Information, Figure S4a,b). The purification of the polymers was achieved by simple precipitation in methanol from their solutions. We noticed that 5bc and 5bd were only soluble in 1,1,1,3,3,3-hexafluoro-2propanol (HFIP) or mixtures of HFIP and chloroform but insoluble in other commonly used organic solvents like chloroform, dimethyl sulfoxide, dimethylformamide (DMF), dimethylacetamide, and tetrahydrofuran. The other polymers were all soluble in chloroform. It was also noted that the indole-based polyester solution in chloroform slowly developed pink coloration at room temperature after 5 days. A thin film of 5ad is shown in Figure S4c, which was prepared using

Figure 1. Previously synthesized indole-based polyesters (a) with a mixture of aliphatic and aromatic ester bonds and new indole-based aromatic polyesters in this work (b).

in PET. Such aliphatic ester bonds are expected to lower the thermal stability of the resulting polyesters. To achieve better thermal properties, indole-based polyesters without labile aliphatic ester bonds are of great interest. Herein, we report our recent synthesis and characterization of two indole-based dicarboxylate monomers with only aromatic ester bonds (i.e., ester bonds derived from aromatic acids) and their polymerization with five bio-based aliphatic diols to yield in total 10 indole-based polyesters. The thermal, mechanical, and rheological properties of the resulting polyesters were reported. According to our result, the absence of weak aliphatic ester

Scheme 1. Synthesis of Monomers 3ab and Polymerization of 5xy Biopolyesters

B

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 2. 1H NMR spectra of monomer 3a and its resulting polyesters 5ay (y = a−e).

Table 1. Summary of Molecular Weight, Thermal and Thermomechanical Analyses, as well as WLF Parameter of Biopolyesters 5xy Tg (°C) sample 5aa 5ab 5ac 5ad 5ae 5ba 5bb 5bc 5bd 5be

Mn (g/mol) 21 300 11 500 35 000 21 300 23 700 14 900 6700 33 500 29 500 17 500

Mw (g/mol) 40 300 20 400 59 800 42 100 41 700 26 800 12 000 60 300 57 300 27 800

PDI 1.89 1.77 1.71 1.98 1.76 1.80 1.80 1.80 2.00 1.60

a

DSC

DMAb

113 94 90 83 71 100 85 87 80 65

125

T1%c 347 302 347 335 330 320 271 330 358 331

104 96 83 112 97 91 76

T5%c 376 354 375 361 360 353 337 365 374 356

Tdc 410 395 419 394 403 390 384 403 405 390

C1d

C2d (K)

Cg1e

Cg2e (K)

9.95

308.3

15.23

201.3

4.80 4.28

205.9 198.3

13.00 13.85

79.5 61.3

3.41

151.8

16.26

31.8

4.53 3.49

205.9 183.1

12.80 14.83

72.9 43.1

DSC: second heat ramp; N2, 10 °C/min. Tg determined as the inflection point of step transition. bDMA: bending mode, 3 °C/min, 1 Hz. Tg determined as the peak maximum in tan δ curves. cTGA: N2, 10 °C/min. dWLF parameters based on reference temperature Tr = 220 °C. eWLF parameters based on the Tg. a

the solution of 5ad in chloroform that was stored at room temperature for 5 days. Polymer films that were prepared directly after the synthesis of the polymers were colorless and stable when being stored at room temperature. The chemical structures of the monomers and polyesters were confirmed by their 1H NMR spectra. Figure 2 shows the 1 H NMR spectra of the polymer series 5ay (all of the polyesters prepared from monomer 3a). For monomer 3a, the chemical shifts of CH2 (a and b) appeared at 2.46 and 4.24 ppm, respectively. The chemical shift of the methyl carboxylate group (c) was found at 3.92 ppm, which was used to monitor the progress of polymerization during the first stage (transesterification). The chemical shifts for the indole units (e−h) were observed at 7.14−8.21 ppm. After the polymerization, all of the signals became broad

(Figures 2 and S2). The signal b (CH2 next to the nitrogen in 3a) for all of the polymers shifted upfield (from 4.10 to 3.90 ppm). Moreover, it was observed that when longer diols were used, the chemical shift of signal b moved closer to the chemical shift of the corresponding peak in the monomer. The same effect was also observed for signal a, which appeared at 2.46 ppm in the monomer spectrum, but moved to higher field for the polymers (2.42−2.29 ppm). The signals for the diol units (c−f) were also clearly observed in all of the polymer spectra. The peak corresponded to the CH2 next to the ester bonds (signal c) was observed at 4.53−4.30 ppm with a downfield shift trend for the polymers with shorter diols. Between 2.20 and 1.40 ppm, the signals corresponded to the other methylene groups (d−f) were observed. According to the 1H NMR spectra, polymers 5ay (y = a−e) were formed. C

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

The 1H NMR spectra of the other series 5by (y = a−e) were also unambiguously assigned (Supporting Information, Figure S3), which indicated the successful formation of the polymers. The molecular weight of the obtained polyesters was determined by size exclusion chromatography (SEC, Table 1). Except for two polymers 5ab and 5bb, all of the other polyesters achieved relatively high molecular weights (Mn ∼ 15 000−35 000 g/mol and Mw ∼ 27 000−60 000 g/mol). For 5ab and 5bb, we have investigated reaction temperature and time, but the resulting polyesters 5ab and 5bb always showed considerably lower molecular weights compared with other polymers. This could be attributed to the lower thermal stability of 5ab and 5bb compared with the other polyesters according to thermogravimetric analysis (TGA) measurements (Table 1). The thermal stability of the obtained polyesters was assessed by TGA (Figure 3). A single-step weight-loss profile was

between the T1% values and the corresponding weight average molecular weight (Mw; see the inset of Figure 3). It is unclear whether the lower molecular weight was caused by the decreased thermal stability of the polyesters or the lower thermal stability was caused by the increased content of vulnerable chain ends for lower-molecular-weight polyesters. Thermal Characterization. The thermal transition of the obtained indole-based polyesters was studied by differential scanning calorimetry (DSC) (Figure 4). All of the obtained polyesters were completely amorphous without any melting endotherm on the heating curves. The glass transitions were clearly observed with Tg values in the range of 71−113 °C for 5ay series and 65−100 °C for 5by series. As shown in Figure 5, the Tg values of the same series of

Figure 5. Tg values of polyester series 5ay (red) and series 5by (blue) measured by DSC (a) and DMA (b) as a function of the number of CH2 groups in the diols used for the synthesis. Lines serve to guide the eye.

Figure 3. TGA weight-loss curves of indole-based polyesters 5xy (x = a−e, y = a−e). The inset shows the correlation between the weight average molecular weight (Mw) and the temperature for a 1% weight loss (T1%).

polyesters (5ay and 5by) decreased with the decreased length of the diols used, which was expected because of the higher flexibility of the polymers with longer methylene “bridges”. In addition, 5ay displayed higher Tg values than the corresponding 5by (with the same diol used), due to their higher rigidity imparted by the shorter methylene bridges inside monomer 3a compared to that of monomer 3b. The Tg values were also independently measured by dynamic mechanical analysis (DMA). As shown in Figure 6, the Tg values of the polyesters were determined using peak maximum in the tan δ.

observed (maximum decomposition rate temperature Td > 380 °C) for all of the polyesters, and the onset of weight loss (T5%, temperature for a 5% weight loss) was higher than 330 °C. This showed good thermal stability of the resulting polyesters. A closer examination of the thermograms in the initial stages of the thermal decomposition revealed that the T1% values were all above 270 °C, and there was a significant linear correlation

Figure 4. DSC second heating curves of polyesters 5ay (a) and 5by (b), including Tg values of polylactic acid (PLA), PET, and Akestra.34,70,71 D

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 6. Storage modulus E′ and tan δ as a function of temperature for series 5ay (a) and series 5by (b).

Polyesters 5ab and 5bb could not be measured by DMA due to the insufficient physical integrity for the samples prepared for DMA measurements, which was possibly caused by their relatively low molecular weights. In fact, the effects of the relatively low molecular weights of 5ab and 5bb were also reflected in their lower Tg values with respect to the expected values (Figure 5a). Expectedly, the Tg values measured by DMA showed the same trend as those obtained from the DSC (Figure 5). The differences in the Tg values according to DSC and DMA are typical, which arise from the different processes that were measured.72 In addition, DMA also provided valuable information regarding the mechanical characteristics of the obtained polyesters in the solid state. Figure 6 shows the storage modulus (E′) and the tan δ as a function of the temperature. All of the polyesters displayed typical temperature-unaffected glassy modulus below Tg. The observed E′ values at 30 °C were found between 1.8 and 2.7 GPa for 5ay and 1.7 and 2.2 GPa for 5by. Both series showed decreasing E′ values as longer diols were used. Rheological Analysis. Polyesters having a Tg superior to PET (76 °C), namely, 5aa, 5ac, 5ad, 5ba, 5bc, and 5bd, were studied by melt rheology.34 The time−temperature superposition was applied to the data obtained from the measurements in a frequency range between 500 and 0.1 rad/s at 10 °C intervals from 170 to 220 °C (5aa was studied between 180 and 230 °C). The thermorheological simplicity of these polyesters was confirmed as good overlap was attained by shifting moduli and viscosity into master curves. A reference temperature (Tr ∼ 220 °C) was used for the comparison among all of the obtained polyesters in a frequency range of five decades up to 104 rad/s. As a representative plot, Figure 7 shows the master curve of both storage (G′) and loss (G″) moduli of 5bd, which depicts its typical linear viscoelastic behavior at the terminal zone and beyond the limit of the rubbery plateau region. At the crossover frequency ωt (where G′ = G″), a typical terminal relaxation time τt = 1/ωt was observed. This relaxation was consistent with the disentanglement relaxation time, which was observed in all of the measured polyesters (Figure S5). Furthermore, the storage modulus G′ (Figure 7) showed an additional relaxation time at low frequencies, which could be attributed to the observed bubbles in the samples after the rheology measurements. The deformation of these bubbles during the oscillatory shear deformation can cause an additional elastic component that dominates at low frequencies.73,74 As a result, an increased

Figure 7. Storage G′ and loss G″ moduli master curves of 5bd at Tr = 220 °C.

G′ but unchanged G″ was observed by the slopes of 0.95−0.97, which was consistent with the expected slope of G″ ≈ 1 for non-cross-linked polymer melts at low frequencies in the terminal region. Figure 8 shows the complex viscosity η* master curves of polyesters 5aa, 5ac, 5ad, 5ba, 5bc, and 5bd. In the low-

Figure 8. Complex viscosity η* master curves of 5aa, 5ac, 5ad, 5ba, 5bc, and 5bd at Tr = 220 °C. E

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Figure 9. WLF fittings of temperature shift factor aT as a function of temperature using a reference temperature of 220 °C (a), and the WLF parameters of indole-based aromatic polyesters compared with that of commercial polymers in the literature (b).75−77

frequency limit, the elastic component to η* disappears and the viscosity becomes independent of the frequency as the material behaves increasingly Newtonian. This frequencyindependent viscosity is the zero-shear viscosity η0, which is approximately proportional to Mw for most polymers.75 However, the soft increase in the complex viscosity observed at low frequencies can be ascribed to the increase of G′ caused by the existence of bubbles. At higher frequencies, a typical shear-thinning behavior was observed as the polyesters became non-Newtonian. All of the polyesters displayed comparable η0 values, which was consistent with their decent molecular weight. All of the six measured polyesters showed thermorheological simplicity, which obeyed the Williams−Landel−Ferry (WLF) equation log (a T) =

−C1(T − Tr) C2 + (T − Tr)

obtained polyesters contained only aromatic ester bonds, which enhanced their thermal stability and structural rigidity. Varied lengths of methylene bridges were contained in the new indole-based polyesters, which resulted in tunable Tg values in the range of 65−113 °C. This indicated that these polymers could be potentially used in a wide range of applications replacing commercial polyesters like PET, poly(ethylene terephthalate) glycol (PETG), Akestra, or Tritan. The new indole-based polyesters were amorphous and optically transparent, which could be conveniently prepared into nonbrittle thin films. Melt rheology measurements of the new polyesters showed a distinct relaxation associated with disentanglements. In the future, investigations on the copolymerization of indolebased dicarboxylate monomers with rigid bis-phenolic building blocks (e.g., various “phytomonomers” derived from lignin)21 to produce whole aromatic polyesters may have the potential to further improve the thermal and mechanical properties of indole-based polyesters toward high-performance engineering bioplastics.

(1)



In the WLF equation, aT is the temperature shift factor employed to generate the master curves, T is the temperature, Tr is an arbitrary reference temperature (220 °C in this study), and C1 and C2 are the empirical constants obtained from curvefitting.75 As shown in Figure 9b, the WLF fits displayed a typical behavior for aT against the temperature. The values of C1 and C2 depended on Tr, and they can be transformed into Cg1 and Cg2 values according to eqs 2 and 3 below that are comparable with literature values.75 As shown in Table 1 and Figure 9b, the Cg1 and Cg2 values for the obtained polyesters 5ac, 5ad, 5ba, 5bc, and 5bd were consistent with the literature values for commercial polymers like PET, polylactic acid, etc.75−77 Interestingly, the most rigid polyester that we prepared (5aa) showed a significantly deviated Cg2 value, for which the exact reason still remained to be unraveled. C1C2 C2 + (Tg − Tr)

(2)

C2g = C2 + (Tg − Tr)

(3)

C1g =



EXPERIMENTAL SECTION Materials. 1,1,1,3,3,3-Hexafluoro-2-propanol (HFIP), 1,3dibromopropane (99%), 1,4-dibromobutane (99%), dibutyltin(IV) oxide DBTO (98%), methyl indole-3-carboxylate (99%), 1,3-propanediol (98%), 1,4-butanediol (>99%), 1,5-pentanediol (>97%), 1,6-hexanediol (>99%), 1,8-octanediol (98%), and sodium hydride (a 60% dispersion in mineral oil) were purchased from Sigma-Aldrich. Toluene (analytical grade, ACS), xylenes (analytical grade, ACS), and chloroform (analytical grade, stabilized with ethanol) were purchased from Scharlau. Methanol was purchased from Honeywell. Dimethylformamide (DMF) (ACS, Reag. Ph. Eur.) was purchased from VWR Chemical. All chemicals were used as received with the exception of DMF, which was obtained from a dry solvent dispenser. Synthesis of Monomers and Polymers. Dimethyl 1,1′(propane-1,3-diyl)bis(1H-indole-3-carboxylate) (Monomer 3a): General Method. To a 300 mL round-bottom flask equipped with a magnetic stirrer, methyl indole-3-carboxylate (20.14 g, 114.9 mmol, 2.10 equiv) and 200 mL dry DMF were added and cooled in an ice bath. To the resulting solution, sodium hydride (5.04 g of a 60% dispersion in mineral oil, 125.8 mmol, 2.30 equiv) was added portionwise and stirred under a N2 blanket. After 60 min, 1,3-dibromopropane (5.56

CONCLUSIONS A series of indole-based polyesters were successfully synthesized by the bulk polycondensation of indole-based dicarboxylate monomers and five aliphatic bio-based diols. The F

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

136.34, 133.88, 126.87, 123.15, 122.25, 121.95, 109.84, 107.80, 63.64, 43.89, 29.73, and 26.04. FT-IR ν (cm−1): 2948, 1688, 1532, 1227, 1198, 1166, 1154, 1098, 1014, 773, and 746. 5ac. White fiberlike powder (95% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (d, 2H, (Hj)), 7.70 (s, 2H, (Hi)), 7.26−7.19 (m, 4H, (Hg, Hh)), 7.12 (d, 2H, (Hf)), 4.37 (t, 4H, (He)), 4.01 (q, 4H, (Hd)), 2.34 (q, 2H, (Hc)), 1.90 (q, 4H, (Hb)), and 1.67 (m, 4H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.02, 136.34, 133.79, 126.85, 123.14, 122.25, 122.02, 122.05, 109.80, 107.97, 63.77, 43.83, 29.77, 28.76, and 23.03. FT-IR ν (cm−1): 2946, 1690, 1533, 1227, 1200, 1166, 1156, 1099, 1014, 775, and 746. 5ae. White fiberlike powder (95% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (d, 2H, (Hk)), 7.73 (s, 2H, (Hj)), 7.28−7.19 (m, 4H, (Hi, Hh)), 7.14 (d, 2H, (Hg)), 4.30 (t, 4H, (Hf)), 4.09 (q, 4H, (He)), 2.42 (q, 2H, (Hd)), 1.78 (q, 4H, (Hc)), and 1.47−1.40 (m, 4H, (Hb, Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.14, 136.41, 133.77, 126.89, 123.20, 122.26, 122.13, 109.82, 108.22, 64.08, 43.93, 29.93, 29.32, 29.05, and 26.19. FT-IR ν (cm−1): 2932, 1683, 1532, 1227, 1199, 1155, 1095, 1013, 773, and 746. 5ba. White fiberlike powder (94% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.17 (d, 2H, (Hi)), 7.64 (s, 2H, (Hh)), 7.23−7.19 (m, 6H, (He, Hf, Hg)), 4.50 (br, 4H, (Hd)), 3.88 (br, 4H, (Hc)), 2.26 (br, 2H, (Hb)), and 1.69 (br, 4H, (Ha)). 13 C NMR (100.61 MHz, CDCl3). δ, ppm: 164.94, 136.41, 134.23, 126.84, 123.04, 122.18, 121.93, 109.93, 107.32, 61.04, 46.32, 28.75, and 27.25. FT-IR ν (cm−1): 3115, 2944, 1690, 1531, 1397, 1269, 1220, 1154, 1089, 1044, 775, and 747. 5bb. White powder (94% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.16 (d, 2H, (Hi)), 7.70 (s, 2H, (Hh)), 7.23− 7.21 (m, 6H, (He, Hf, Hg)), 4.40 (br, 4H, (Hd)), 3.97 (br, 4H, (Hc)), 1.97 (br, 4H, (Hb)), and 1.76 (br, 4H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.07, 136.43, 134.17, 126.83, 122.99, 122.11, 121.93, 109.95, 107.47, 63.59, 46.37, 27.27, and 26.06. FT-IR ν (cm−1): 3109, 2945, 1687, 1531, 1397, 1263, 1220, 1154, 1089, 1014, 775, and 745. 5bc. White spherical-like powder (96% yield). 1H NMR (400.13 MHz, HIFP-d2/CDCl3 (50/50 v/v)) δ, ppm: 8.10 (d, 2H, (Hj)), 7.83 (s, 2H, (Hi)), 7.33−7.26 (m, 6H, (Hf, Hg, Hh)), 4.40 (m, 4H, (He)), 4.14 (br, 4H, (Hd)), 1.98 (br, 4H, (Hc)), 1.89 (br, 4H, (Hb)), and 1.75 (br, 2H, (Ha)). 13C NMR (100.61 MHz, HIFP-d2/CDCl3 (50/50 v/v)). δ, ppm: 168.86, 137.10, 135.90, 126.78, 123.78, 122.76, 121.81, 110.77, 106.81, 65.34, 46.92, 28.63, 27.43, and 22.96. FT-IR ν (cm−1): 3114, 2946, 1691, 1532, 1398, 1265, 1221, 1159, 1092, 1014, 775, and 745. 5bd. White fiberlike powder (93% yield). 1H NMR (400.13 MHz, HIFP-d2/CDCl3 (50/50 v/v)) δ, ppm: 7.84 (br, 2H, (Hj)), 7.35 (s, 2H, (Hi)), 7.35−7.26 (br, 6H, (Hf, Hg, Hh)), 4.39 (br, 4H, (He)), 4.16 (br, 4H, (Hd)), 1.93 (br, 4H, (Hc, Hb)), and 1.65 (br, 8H, (Ha)). 13C NMR (100.61 MHz, HIFP-d2/CDCl3 (50/50 v/v)). δ, ppm: 168.94, 137.11, 135.89, 126.79, 123.79, 122.95, 121.86, 110.79, 106.90, 65.60, 46.93, 28.89, 27.45, and 26.08. FT-IR ν (cm−1): 3114, 2935, 1688, 1533, 1397, 1265, 1221, 1156, 1091, 1014, 775, and 745. 5be. White fiberlike powder (94% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.16 (d, 2H, (Hk)), 7.71 (s, 2H, (Hj)), 7.26−7.24 (m, 6H, (Hi, Hh, Hg)), 4.30 (t, 4H, (Hf)), 4.04 (br, 4H, (He)), 1.82−1.76 (br, 8H, (Hd, Hc)), and 1.47−1.40 (m, 4H, (Hb,Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.25, 136.46, 134.08, 126.83, 123.00, 122.08, 109.91, 107.82,

mL, 54.7 mmol, 1 equiv) was added dropwise. The mixture was stirred for 2 h and allowed to reach room temperature overnight. The mixture was poured into a 1700 mL ice−water mixture and stirred for 2 h. The white solid was filtrated and washed extensively with distilled water. Recrystallization from toluene (×2) afforded 17.86 g of white platelike crystals (yield: 83%). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.21 (d, 2H, (Hh)), 7.74 (s, 2H, (Hg)), 7.29 (td, 2H, (He)), 7.25 (td, 2H, (Hf)), 7.17 (d, 2H, (Hd)), 4.12 (t, 4H, (Hc)), 3.92 (s, 6H, (Hb)), and 2.46 (q, 2H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.34, 136.33, 133.77, 126.86, 123.22, 122.27, 122.07, 109.79, 107.84, 51.16, 43.86, and 29.74. FT-IR ν (cm−1): 3108, 3045, 2950, 1680, 1531, 1265, 1232, 1183, 1153, 1106, 1092, 1013, 775, 749, 728, and 557. HRMS (ESI +) exact mass calcd for C23H22N2O4 391.1658, found 391.1660. mp: 184 °C (DSC). Dimethyl 1,1′-(butane-1,4-diyl)bis(1H-indole-3-carboxylate) (Monomer 3b). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (m, 2H, (Hh)), 7.73 (s, 2H, (Hg)), 7.30−7.27 (m, 3H, (Hd, He, Hf)), 4.10 (q, 4H, (Hc)), 3.9q (s, 6H, (Hb)), and 1.87 (q, 2H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 165.47, 136.42, 134.05, 126.84, 123.06, 122.13, 122.05, 109.87, 107.50, 51.14, 46.50, and 27.34. FT-IR ν (cm−1): 3118, 2946, 1686, 1533, 1404, 1269, 1228, 1149, 1118, 1091, 1028, 775, and 739. HRMS (ESI+) exact mass calcd for C24H24N2O4 405.1819, found 405.1814. mp: 205 °C (DSC). Synthesis of Polyesters. The procedure for the synthesis of 5ad was described as a general protocol, and all of the other polyesters were prepared similarly. 1,6-Hexanediol (0.94 g, 7.94 mmol, 3.0 equiv), monomer 1 (1.03 g, 2.65 mmol, 1 equiv), DBTO (6.59 mg, 1 mol % with respect to 1), and 2 mL xylenes were added to a two-necked round-bottom flask equipped with a mechanical stirrer. After 15 min under reflux and under a flow of N2, xylenes were distilled off and the reaction mixture was heated up to 180 °C. Once transesterification of 1 was completed (approximately 3 h, observed by 1H NMR), the flow of N2 was increased and the mixture was heated to 220 °C for 20 h. The resulting material was dissolved in 7 mL chloroform and precipitated in 500 mL MeOH. The white small fibers formed were filtered off and washed repeatedly with methanol. Drying at 50 °C under vacuum overnight afforded 1.09 g (93%) of a white fiberlike powder. 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (d, 2H, (Hj)), 7.70 (s, 2H, (Hi)), 7.26−7.19 (m, 4H, (Hg, Hh)), 7.12 (d, 2H, (Hf)), 4.37 (t, 4H, (He)), 4.01 (t, 4H, (Hd)), 2.34 (q, 2H, (Hc)), 1.90 (q, 4H, (Hb)), and 1.67 (m, 4H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 164.95, 136.25, 133.66, 126.76, 123.05, 122.13, 121.92, 109.71, 107.92, 63.76, 43.77, 29.73, 28.81, and 25.85. FT-IR ν (cm−1): 2924, 1690, 1533, 1227, 1200, 1164, 1154, 1099, 1038, 1011, 775, and 747. 5aa. White fiberlike powder (92% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (d, 2H, (Hi)), 7.63 (s, 2H, (Hh)), 7.24−7.18 (m, 4H, (Hf, Hg)), 7.06 (d, 2H, (He)), 4.53 (t, 4H, (Hd)), 3.88 (br, 4H, (Hc)), 2.29 (m, 2H, (Hb)), and 2.21 (m, 4H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 164.81, 136.29, 133.96, 126.85, 123.19, 122.30, 121.98, 109.82, 107.65, 61.09, 43.78, 29.67, and 28.72. FT-IR ν (cm−1): 2924, 1690, 1533, 1227, 1200, 1164, 1154, 1099, 1038, 1011, 775, and 747. 5ab. White powder (99% yield). 1H NMR (400.13 MHz, CDCl3) δ, ppm: 8.18 (d, 2H, (Hi)), 7.70 (s, 2H, (Hh)), 7.25− 7.19 (m, 4H, (Hf, Hg)), 7.10 (d, 2H, (He)), 4.42 (br, 4H, (Hd)), 4.02 (br, 4H, (Hc)), 2.34 (br, 2H, (Hb)), and 1.99 (br, 4H, (Ha)). 13C NMR (100.61 MHz, CDCl3). δ, ppm: 164.94, G

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

64.02, 46.47, 29.32, 29.06, 27.41, and 26.19. FT-IR ν (cm−1): 2927, 1683, 1531, 1223, 1198, 1154, 1089, 1013, 775, and 744. Analytical Methods. Nuclear magnetic resonance (NMR) measurements were carried out on a Bruker DRX400 spectrometer at a proton frequency of 400.13 MHz and a carbon frequency of 100.61 MHz. Fourier transform infrared (FT-IR) spectra were measured with an attenuated total reflection setup using a Bruker Alpha FT-IR spectrometer. Twenty-four scans were coadded using a resolution of 4 cm−1. Size exclusion chromatography (SEC) was performed in chloroform at 35 °C with a flow rate of 1.0 mL/min. The SEC equipment used was a Viscotek 305 TDA, which included a guard column and two Malvern Panalytical general purpose mixed bed columns with an exclusion limit of 20 × 106 Da for polystyrene. Detection consisted of a conventional dual-cell refractive index detector, a four-capillary bridge viscometer, and a light scattering detector operating at 3 mW, at a wavelength of 670 nm, and measurement angles of 90 and 7o. The three detectors were calibrated with a polystyrene standard (96 kDa) from Polymers Laboratories. Molecular weights were determined by the triple-detection method using the OmiSEC. 5.12. Software (Malvern). Differential scanning calorimetry (DSC) measurements were performed using TA Instruments DSC Q2000. The samples were studied with a heating rate of 10 °C/min under nitrogen with a purge rate of 50 mL/min. The sequence consisted of a heating ramp from 40 to 200 °C, followed by a cooling ramp to −50 °C, and finally a heating ramp to 220 °C, which was employed to determine the glass-transition temperature (Tg). Thermogravimetric analysis (TGA) was performed with a thermogravimetric analyzer TA Instruments Q500 at a heating rate of 10 °C/min, under nitrogen with a purge rate of 50 mL/min. Dynamic mechanical analysis (DMA) was performed on a TA Instruments Q800 dynamic mechanical analyzer in an oscillatory bending mode at a frequency of 1 Hz, an oscillatory amplitude of 15 μm, and with a temperature ramp of 3 °C. Optically clear freestanding rectangular specimens (l: 8 mm, w: 5 mm, t: 1 mm) were prepared by compression-molding at 200 °C for 10 min. The rheological analysis was performed on an Advance Rheometer AR2000 ETC from TA Instruments. The characterization was carried out under a N2 atmosphere and using a disposable 20 mm diameter aluminum parallel-plate geometry with a gap of ca. 1 mm. Dynamic strain sweep measurements were carried out at 1 Hz from a 0.1 to 100% oscillatory strain to determine the viscoelastic linear region. Dynamic frequency sweep experiments were performed in the linear viscoelastic region at an oscillatory strain of 2% and a frequency range of 0.1−500 rad/s.



ORCID

Baozhong Zhang: 0000-0002-7308-1572 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the Åforsk Foundation (No. 16-479), Mistra Foundation (the “STEPS” project, No. 2016/1489), the Crafoord Foundation (Nos. 20160774 and 20180939), EIT Climate-KIC flagship eCircular, and the Royal Physiographic Society in Lund.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01802.



REFERENCES

(1) Jambeck, J. R.; Geyer, R.; Wilcox, C.; Siegler, T. R.; Perryman, M.; Andrady, A.; Narayan, R.; Law, K. L. Marine Pollution. Plastic Waste Inputs from Land into the Ocean. Science 2015, 347, 768−771. (2) Frost, J. W.; Miermont, A.; Schweitzer, D.; Bui, V.; Paschke, E.; Wicks, D. A. Biobased Polyesters. U.S. Patent US8,415,496B2, 2013. (3) Miller, K. K.; Zhang, P.; Nishizawa-Brennen, Y.; Frost, J. W. Synthesis of Biobased Terephthalic Acid from Cycloaddition of Isoprene with Acrylic Acid. ACS Sustainable Chem. Eng. 2014, 2, 2053−2056. (4) Banella, M. B.; Gioia, C.; Vannini, M.; Colonna, M.; Celli, A.; Gandini, A. A Sustainable Route to a Terephthalic Acid Precursor. ChemSusChem 2016, 9, 942−945. (5) Bérard, S.; Vallée, C.; Delcroix, D. Sorbic Acid as a Renewable Resource for Atom-Economic and Selective Production of p-Toluic Acid and Alkyl-p-Toluates: Intermediates to Bioterephthalic Acid and Esters. Ind. Eng. Chem. Res. 2015, 54, 7164−7168. (6) Kraus, G. A.; Riley, S.; Cordes, T. Aromatics from Pyrones: ParaSubstituted Alkyl Benzoates from Alkenes, Coumalic Acid and Methyl Coumalate. Green Chem. 2011, 13, 2734. (7) Tachibana, Y.; Kimura, S.; Kasuya, K. Synthesis and Verification of Biobased Terephthalic Acid from Furfural. Sci. Rep. 2015, 5, No. 8249. (8) Williams, C. L.; Chang, C.-C.; Do, P.; Nikbin, N.; Caratzoulas, S.; Vlachos, D. G.; Lobo, R. F.; Fan, W.; Dauenhauer, P. J. Cycloaddition of Biomass-Derived Furans for Catalytic Production of Renewable p-Xylene. ACS Catal. 2012, 2, 935−939. (9) Chang, C.-C.; Green, S. K.; Williams, C. L.; Dauenhauer, P. J.; Fan, W. Ultra-Selective Cycloaddition of Dimethylfuran for Renewable p-Xylene with H-BEA. Green Chem. 2014, 16, 585−588. (10) Lin, Z.; Ierapetritou, M.; Nikolakis, V. Aromatics from Lignocellulosic Biomass: Economic Analysis of the Production of pXylene from 5-Hydroxymethylfurfural. AIChE J. 2013, 59, 2079− 2087. (11) Pacheco, J. J.; Labinger, J. A.; Sessions, A. L.; Davis, M. E. Route to Renewable PET: Reaction Pathways and Energetics of Diels−Alder and Dehydrative Aromatization Reactions Between Ethylene and Biomass-Derived Furans Catalyzed by Lewis Acid Molecular Sieves. ACS Catal. 2015, 5, 5904−5913. (12) Colonna, M.; Berti, C.; Fiorini, M.; Binassi, E.; Mazzacurati, M.; Vannini, M.; Karanam, S. Synthesis and Radiocarbon Evidence of Terephthalate Polyesters Completely Prepared from Renewable Resources. Green Chem. 2011, 13, 2543. (13) Mialon, L.; Pemba, A. G.; Miller, S. A. Biorenewable Polyethylene Terephthalate Mimics Derived from Lignin and Acetic Acid. Green Chem. 2010, 12, 1704. (14) Mialon, L.; Vanderhenst, R.; Pemba, A. G.; Miller, S. A. Polyalkylenehydroxybenzoates (PAHBs): Biorenewable Aromatic/ Aliphatic Polyesters from Lignin. Macromol. Rapid Commun. 2011, 32, 1386−1392. (15) Pion, F.; Ducrot, P.-H.; Allais, F. Renewable Alternating Aliphatic-Aromatic Copolyesters Derived from Biobased Ferulic Acid, Diols, and Diacids: Sustainable Polymers with Tunable Thermal Properties. Macromol. Chem. Phys. 2014, 215, 431−439.

Synthetic pathway toward methyl indole-3-carboxylate; zoom-in 1H NMR spectra; FT-IR spectra; storage G′ and loss G″ moduli master curves; DSC traces of monomers; and photographs of polymer samples (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. H

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

on Effect of Catalyst Type on Molecular Weight Increase. Polymers 2019, 11, No. 438. (34) Burgess, S. K.; Leisen, J. E.; Kraftschik, B. E.; Mubarak, C. R.; Kriegel, R. M.; Koros, W. J. Chain Mobility, Thermal, and Mechanical Properties of Poly(Ethylene Furanoate) Compared to Poly(Ethylene Terephthalate). Macromolecules 2014, 47, 1383−1391. (35) Rosenboom, J.-G.; Hohl, D. K.; Fleckenstein, P.; Storti, G.; Morbidelli, M. Bottle-Grade Polyethylene Furanoate from RingOpening Polymerisation of Cyclic Oligomers. Nat. Commun. 2018, 9, No. 2701. (36) Ma, Q.; Qu, Y.; Zhang, X.; Liu, Z.; Li, H.; Zhang, Z.; Wang, J.; Shen, W.; Zhou, J. Systematic Investigation and Microbial Community Profile of Indole Degradation Processes in Two Aerobic Activated Sludge Systems. Sci. Rep. 2016, 5, No. 17674. (37) Bandini, M.; Eichholzer, A. Catalytic Functionalization of Indoles in a New Dimension. Angew. Chem., Int. Ed. 2009, 48, 9608− 9644. (38) Barden, T. C. Indoles: Industrial, Agricultural and Over-theCounter Uses. In Heterocyclic Scaffolds II: Reactions and Applications of Indoles; Gribble, G. W., Ed.; Springer: Berlin, Heidelberg, 2010; pp 31−46. (39) Sundberg, R. J. Indole. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons, Inc: Hoboken, NJ, 2000. (40) Li, X.; Baryshnikov, G.; Deng, C.; Bao, X.; Wu, B.; Zhou, Y.; Ågren, H.; Zhu, L. A Three-Dimensional Ratiometric Sensing Strategy on Unimolecular Fluorescence−thermally Activated Delayed Fluorescence Dual Emission. Nat. Commun. 2019, 10, No. 731. (41) Gribble, G. W. Recent Developments in Indole Ring Synthesismethodology and Applications. J. Chem. Soc. Perkin Trans. 1 2000, 0, 1045−1075. (42) Yao, Q.; Xu, L.; Han, Z.; Zhang, Y. Production of Indoles via Thermo-Catalytic Conversion and Ammonization of Bio-Derived Furfural. Chem. Eng. J. 2015, 280, 74−81. (43) Splitstoser, J. C.; Dillehay, T. D.; Wouters, J.; Claro, A. Early Pre-Hispanic Use of Indigo Blue in Peru. Sci. Adv. 2016, 2, No. e1501623. (44) Llabres-Campaner, P. J.; Ballesteros-Garrido, R.; Ballesteros, R.; Abarca, B. Straight Access to Indoles from Anilines and Ethylene Glycol by Heterogeneous Acceptorless Dehydrogenative Condensation. J. Org. Chem. 2018, 83, 521−526. (45) Yao, Q.; Xu, L.; Zhang, Y.; Fu, Y. Enhancement of Indoles Production and Catalyst Stability in Thermo-Catalytic Conversion and Ammonization of Furfural with NH3 and N2 Environments. J. Anal. Appl. Pyrolysis 2016, 121, 258−266. (46) Xu, L.; Yao, Q.; Deng, J.; Han, Z.; Zhang, Y.; Fu, Y.; Huber, G. W.; Guo, Q. Renewable N-Heterocycles Production by Thermocatalytic Conversion and Ammonization of Biomass over ZSM-5. ACS Sustainable Chem. Eng. 2015, 3, 2890−2899. (47) Xu, L.; Jiang, Y.; Yao, Q.; Han, Z.; Zhang, Y.; Fu, Y.; Guo, Q.; Huber, G. W. Direct Production of Indoles via Thermo-Catalytic Conversion of Bio-Derived Furans with Ammonia over Zeolites. Green Chem. 2015, 17, 1281−1290. (48) Huang, F.; Tahmasebi, A.; Maliutina, K.; Yu, J. Formation of Nitrogen-Containing Compounds during Microwave Pyrolysis of Microalgae: Product Distribution and Reaction Pathways. Bioresour. Technol. 2017, 245, 1067−1074. (49) Han, T. H.; Lee, J.-H.; Cho, M. H.; Wood, T. K.; Lee, J. Environmental Factors Affecting Indole Production in Escherichia coli. Res. Microbiol. 2011, 162, 108−116. (50) Arevalo-Villena, M.; Bartowsky, E. J.; Capone, D.; Sefton, M. A. Production of Indole by Wine-Associated Microorganisms under Oenological Conditions. Food Microbiol. 2010, 27, 685−690. (51) Ma, Q.; Zhang, X.; Qu, Y. Biodegradation and Biotransformation of Indole: Advances and Perspectives. Front. Microbiol. 2018, 9, No. 2625. (52) Kim, H.-J.; Jang, S.; Kim, J.; Yang, Y.-H.; Kim, Y.-G.; Kim, B.G.; Choi, K.-Y. Biosynthesis of Indigo in Escherichia coli Expressing Self-Sufficient CYP102A from Streptomyces cattleya. Dyes Pigm. 2017, 140, 29−35.

(16) Gioia, C.; Banella, M. B.; Marchese, P.; Vannini, M.; Colonna, M.; Celli, A. Advances in the Synthesis of Bio-Based Aromatic Polyesters: Novel Copolymers Derived from Vanillic Acid and εCaprolactone. Polym. Chem. 2016, 7, 5396−5406. (17) Gioia, C.; Banella, M. B.; Totaro, G.; Vannini, M.; Marchese, P.; Colonna, M.; Sisti, L.; Celli, A. Biobased Vanillic Acid and Ricinoleic Acid: Building Blocks for Fully Renewable Copolyesters. J. Renewable Mater. 2018, 6, 126−135. (18) Fache, M.; Boutevin, B.; Caillol, S. Vanillin, a Key-Intermediate of Biobased Polymers. Eur. Polym. J. 2015, 68, 488−502. (19) Wilsens, C. H. R. M.; Verhoeven, J. M. G. A.; Noordover, B. A. J.; Hansen, M. R.; Auhl, D.; Rastogi, S. Thermotropic Polyesters from 2,5-Furandicarboxylic Acid and Vanillic Acid: Synthesis, Thermal Properties, Melt Behavior, and Mechanical Performance. Macromolecules 2014, 47, 3306−3316. (20) Gioia, C.; Banella, M. B.; Vannini, M.; Celli, A.; Colonna, M.; Caretti, D. Resorcinol: A Potentially Bio-Based Building Block for the Preparation of Sustainable Polyesters. Eur. Polym. J. 2015, 73, 38−49. (21) Kaneko, T.; Thi, T. H.; Shi, D. J.; Akashi, M. Environmentally Degradable, High-Performance Thermoplastics from Phenolic Phytomonomers. Nat. Mater. 2006, 5, 966−970. (22) Nguyen, H. T. H.; Reis, M. H.; Qi, P.; Miller, S. A. Polyethylene Ferulate (PEF) and Congeners: Polystyrene Mimics Derived from Biorenewable Aromatics. Green Chem. 2015, 17, 4512− 4517. (23) Nguyen, H. T. H.; Short, G. N.; Qi, P.; Miller, S. A. Copolymerization of Lactones and Bioaromatics via Concurrent RingOpening Polymerization/Polycondensation. Green Chem. 2017, 19, 1877−1888. (24) Fonseca, A. C.; Lima, M. S.; Sousa, A. F.; Silvestre, A. J.; Coelho, J. F. J.; Serra, A. C. Cinnamic Acid Derivatives as Promising Building Blocks for Advanced Polymers: Synthesis, Properties and Applications. Polym. Chem. 2019, 10, 1696−1723. (25) Burgess, S. K.; Karvan, O.; Johnson, J. R.; Kriegel, R. M.; Koros, W. J. Oxygen Sorption and Transport in Amorphous Poly(Ethylene Furanoate). Polymer 2014, 55, 4748−4756. (26) Papageorgiou, G. Z.; Tsanaktsis, V.; Papageorgiou, D. G.; Chrissafis, K.; Exarhopoulos, S.; Bikiaris, D. N. Furan-Based Polyesters from Renewable Resources: Crystallization and Thermal Degradation Behavior of Poly (Hexamethylene 2,5-Furan-Dicarboxylate). Eur. Polym. J. 2015, 67, 383−396. (27) 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− 8664. (28) Sousa, 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. (29) Pan, T.; Deng, J.; Xu, Q.; Zuo, Y.; Guo, Q. X.; Fu, Y. Catalytic Conversion of Furfural into a 2,5-Furandicarboxylic Acid-Based Polyester with Total Carbon Utilization. ChemSusChem 2013, 6, 47− 50. (30) 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. (31) Banella, M. B.; Bonucci, J.; Vannini, M.; Marchese, P.; Lorenzetti, C.; Celli, A. Insights into the Synthesis of Poly(Ethylene 2,5-Furandicarboxylate) from 2,5-Furandicarboxylic Acid: Steps toward Environmental and Food Safety Excellence in Packaging Applications. Ind. Eng. Chem. Res. 2019, 58, 8955−8962. (32) Joshi, A. S.; Alipourasiabi, N.; Kim, Y.-W.; Coleman, M. R.; Lawrence, J. G. Role of Enhanced Solubility in Esterification of 2,5Furandicarboxylic Acid with Ethylene Glycol at Reduced Temperatures: Energy Efficient Synthesis of Poly(Ethylene 2,5-Furandicarboxylate). React. Chem. Eng. 2018, 3, 447−453. (33) Chebbi, Y.; Kasmi, N.; Majdoub, M.; Papageorgiou, G. Z.; Achilias, D. S.; Bikiaris, D. N. Solid-State Polymerization of Poly(Ethylene Furanoate) Biobased Polyester, III: Extended Study I

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

Dynamic Mechanical Spectra. J. Therm. Anal. Calorim. 2014, 116, 447−453. (73) Wolff, F.; Münstedt, H. Artefacts of the Storage Modulus Due to Bubbles in Polymeric Fluids. Rheol. Acta 2013, 52, 287−289. (74) Dijkstra, D. J. Guidelines for Rheological Characterization of Polyamide Melts (IUPAC Technical Report). Pure Appl. Chem. 2009, 81, 339−349. (75) Ferry, J. D. Viscoelastic Properties of Polymers, 3rd ed.; John Wiley & Sons: New York, 1980. (76) Mark, J. E. Physical Properties of Polymers Handbook, 2nd ed.; Mark, J. E., Ed.; Springer Science+Business Media, LLC: New York, 2007. (77) Dorgan, J. R.; Janzen, J.; Clayton, M. P.; Hait, S. B.; Knauss, D. M. Melt Rheology of Variable L-Content Poly(Lactic Acid). J. Rheol. 2005, 49, 607−619.

(53) Zhou, W.; Xu, J. Progress in Conjugated Polyindoles: Synthesis, Polymerization Mechanisms, Properties, and Applications. Polym. Rev. 2017, 57, 248−275. (54) Cheng, Y.-J.; Yang, S.-H.; Hsu, C.-S. Synthesis of Conjugated Polymers for Organic Solar Cell Applications. Chem. Rev. 2009, 109, 5868−5923. (55) Chen, Q.; Han, B.-H. Microporous Polycarbazole Materials: From Preparation and Properties to Applications. Macromol. Rapid Commun. 2018, 39, No. 1800040. (56) Yang, P.; Yang, L.; Wang, Y.; Song, L.; Yang, J.; Chang, G. An Indole-Based Aerogel for Enhanced Removal of Heavy Metals from Water via the Synergistic Effects of Complexation and Cation−π Interactions. J. Mater. Chem. A 2019, 7, 531−539. (57) Shen, X.-Y.; Tang, C.-C.; Jan, J.-S. Synthesis and Hydrogelation of Star-Shaped Poly(l-Lysine) Polypeptides Modified with Different Functional Groups. Polymer 2018, 151, 108−116. (58) Liu, J.; Gu, P.-Y.; Li, N.-J.; Wang, L.-H.; Zhang, C.-Y.; Xu, Q.F.; Lu, J.-M. Preparation of a Polymer Containing Indole Groups by RAFT Polymerization and One-Phase Synthesis of AuNPs-Polymer Nanocomposites. J. Appl. Polym. Sci. 2013, 129, 2913−2921. (59) Zhao, C.; Ouyang, K.; Yang, N.; Zhang, J.; Yang, Z. Synthesis and Properties of Optically Active Helical Polyethers Bearing Indole or Carbazole Derivatives. Macromol. Res. 2016, 24, 393−399. (60) Boopathy, M.; Selvam, R.; JohnSanthoshkumar, S.; Subramanian, K. Synthesis and Evaluation of Polyacrylamides Derived from Polycyclic Pendant Naphthalene, Indole, and Phenothiazine Based Chalcone Moiety as Potent Antimicrobial Agents. Polym. Adv. Technol. 2017, 28, 717−727. (61) Grigalevicius, S.; Zostautiene, R.; Sipaviciute, D.; Stulpinaite, B.; Volyniuk, D.; Grazulevicius, J. V.; Liu, L.; Xie, Z.; Zhang, B. Polymers Containing Diphenylvinyl-Substituted Indole Rings as Charge-Transporting Materials for OLEDs. J. Electron. Mater. 2016, 45, 1210−1215. (62) Zuo, P.; Su, Y.; Li, W. Comb-Like Poly(Ether-Sulfone) Membranes Derived from Planar 6,12-Diaryl-5,11-Dihydroindolo[3,2b]Carbazole Monomer for Alkaline Fuel Cells. Macromol. Rapid Commun. 2016, 37, 1748−1753. (63) Wei, W.; Yang, L.; Chang, G. Heat-Resistant and Photoluminescent Indole-Based Poly(Ether Sulfone)S. High Perform. Polym. 2018, 30, 475−479. (64) Weiner, B.-Z.; Zilkha, A. Condensation Polymers Containing Indole Derivatives. J. Macromol. Sci., Chem. 1980, 14, 379−388. (65) Wang, P.; Arza, C. R.; Zhang, B. Indole as a New Sustainable Aromatic Unit for High Quality Biopolyesters. Polym. Chem. 2018, 9, 4706−4710. (66) Yoo, W.-J.; Capdevila, M. G.; Du, X.; Kobayashi, S. BaseMediated Carboxylation of Unprotected Indole Derivatives with Carbon Dioxide. Org. Lett. 2012, 14, 5326−5329. (67) Kasmi, N.; Majdoub, M.; Papageorgiou, G. Z.; Achilias, D. S.; Bikiaris, D. N. Solid-State Polymerization of Poly(Ethylene Furanoate) Biobased Polyester, I: Effect of Catalyst Type on Molecular Weight Increase. Polymers 2017, 9, No. 607. (68) Terzopoulou, Z.; Karakatsianopoulou, E.; Kasmi, N.; Tsanaktsis, V.; Nikolaidis, N.; Kostoglou, M.; Papageorgiou, G. Z.; Lambropoulou, D. A.; Bikiaris, D. N. Effect of Catalyst Type on Molecular Weight Increase and Coloration of Poly(Ethylene Furanoate) Biobased Polyester during Melt Polycondensationin. Polym. Chem. 2017, 8, 6895−6908. (69) Lingier, S.; Spiesschaert, Y.; Dhanis, B.; De Wildeman, S.; Du Prez, F. E. Rigid Polyurethanes, Polyesters, and Polycarbonates from Renewable Ketal Monomers. Macromolecules 2017, 50, 5346−5352. (70) Becker, J. M.; Pounder, R. J.; Dove, A. P. Synthesis of Poly(Lactide)s with Modified Thermal and Mechanical Properties. Macromol. Rapid Commun. 2010, 31, 1923−1937. (71) Hatti-kaul, R.; Nilsson, L. J.; Zhang, B.; Rehnberg, N.; Lundmark, S. Designing Biobased Recyclable Polymers for Plastics. Trends Biotechnol. 2019, 1−18. (72) Lei, Z.; Xing, W.; Wu, J.; Huang, G.; Wang, X.; Zhao, L. The Proper Glass Transition Temperature of Amorphous Polymers on J

DOI: 10.1021/acsomega.9b01802 ACS Omega XXXX, XXX, XXX−XXX