Effects of Various 1,3-Propanediols on the ... - ACS Publications

Jan 2, 2019 - annihilation lifetime spectroscopy (PALS). The substituent ... break is only about 5%, which is much lower than that of PET. As for PPF ...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Effects of Various 1, 3-Propanediols on the Properties of Poly(propylene furandicarboxylate) Jinggang Wang, liyuan Sun, Zhisen Shen, Jin Zhu, Xingliang Song, and Xiaoqing Liu ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b05288 • Publication Date (Web): 02 Jan 2019 Downloaded from http://pubs.acs.org on January 11, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Effects of Various 1, 3-Propanediols on the Properties of Poly(propylene furandicarboxylate)

Jinggang Wang 1, Liyuan Sun 1, 3#, Zhisen Shen 4, Jin Zhu 1, Xingliang Song 2*, Xiaoqing Liu 1 *

1 Key Laboratory of Bio-based Polymeric Materials Technology and Application of Zhejiang Province, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo, 315201 (P. R. China) 2 University of Chinese Academy of Sciences, Beijing 100049 (P. R. China) 3 School of Chemistry and Chemical Engineering, Linyi University, Linyi 276005, China 4 Department of Otorhinolaryngology Head and Neck Surgery, Lihuili Hospital of Ningbo University, Ningbo, Zhejiang 315040, (P. R. China) Corresponding Author: Dr. Xiaoqing Liu, ([email protected]); Dr. Xingliang Song, ([email protected]) # Equal contribution to the first author

1

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT Diols or acids with different skeletal structures could be used to polymerize 2,5furandicarboxylic acid (FDCA) in order to adjust the properties of FDCA-based polyesters. A series of FDCA polyesters with similar skeletal structure as poly(propylene furandicarboxylate) (PPF) were prepared from FDCA and 1,3propanediols containing different substituent groups. The effect of substituent groups on the thermal properties and gas barrier behaviors were investigated by differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), dynamic mechanical analysis (DMA), rheological analysis (RA) and positron annihilation lifetime spectroscopy (PALS). The substituent -CH3 significantly influenced the thermal properties of the polyesters, where the glass transition temperature (Tg) and crystallizability increased from PPF to poly(neopentyl glycol furandicarboxylate) (PNF), and then decreased from PNF to poly(2-ethyl-2-butyl-1,3-propylene furandicarboxylate) (PEBF). PNF displayed the highest Tg of 70 oC and Tm of 201 oC with ΔHm of 32.1 J/g. PPF possessed a Tm of 173 oC with ΔHm of 0.9 J/g, while poly(2methyl-1,3-propylene furandicarboxylate) (PMF) was an amorphous polyester. The gas barrier properties followed the trend of PPF > PMF > PNF due to the increased β relaxation and fractional free volume (FFV) after the introduction of lateral -CH3 groups.

Key Words: 2,5-Furandicarboxylic acid (FDCA); polyesters; lateral groups; barrier properties; free fractional volume.

2

ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

INTRODUCTION In recent years, driven by the increasing concern on environmental issues and rapid crude oil consumption, the bio-based 2,5-furandicarboxylic acid (FDCA) has attracted more attention due to its great potential to replace the petroleum-based terephthalate (TPA). 1 Up to now, especially in the polymer industry, a large number of polyesters derived from FDCA, including poly (ethylene 2,5-furandicarboxylate) (PEF), poly(propylene

2,5-furandicarboxylate)

8-11

(PPF),

poly(butylene

furandicarboxylate) (PBF),12-16 poly(octylene 2,5-furandicarboxylate) (POF), poly(hexamethylene 2,5-furandicarboxylate) (PHF), (PDeF)

21,22,

20

poly(dodecylene 2,5-furanoate) (PDoF)

23

2-7

2,517-19

poly(decylene-2,5-furanote) and poly(neopentyl glycol

furandicarboxylate) (PNF) 24, 25 have been widely investigated. Among them, PEF, PPF and PBF attract greater attention due to the similar mechanical or thermal properties to their petroleum-based counterparts, i.e. poly(ethylene terephthalate) (PET), poly(propylene terephthalate) (PPT)

3

and poly(butylene terephthalate) (PBT).

26

3, 27

Taking PEF as an example, it demonstrates a higher glass transition temperature (Tg), 1.6 times higher tensile modulus, 28, 29 and much better gas barrier property (13-19 and 6-11 times higher barrier to CO2 and O2) when compared to PET.

30, 31

However, its

elongation at break is only about 5%, which is much lower than that of PET. As for PPF and PBF, they demonstrate much weaker crystallizabilities with respect to PPT and PBT,

13,

32

which significantly limits their application. Therefore, the

comprehensive properties of FDCA-based polyesters still have large space for further modification or improvement to meet various demands from industry. That is why there have been attempts to polymerize numerous diols or acids with FDCA. There are two reported strategies to adjust or improve properties of FDCA polyesters. The first one is using diols with varied structures.

2, 8-23

As mentioned above, various

aliphatic diols, from ethylene glycol to dodecylene glycol, have been employed and properties of obtained polyesters are shown in Figure 1. It was noted that their melting temperature (Tm) and tensile strength gradually decreased with the length of diols due to the increased molecular flexibility. The other method is copolymerization with the 3

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

third co-monomers. 15,33-36 So far, a large quantity of organic acids, including succinic acid, 33 adipic acid, 34 terephthalate, 35 lactic acid 36 as well as dimethyl carbonate,15 and various diols, 4,7,37,38 such as 2,5-bis(hydroxymethyl)-furan, 37 isosorbide 38 and 2, 2, 4, 4-tetramethyl-1, 3-cyclobutanediol (CBDO),

4, 7

have been employed as the co-

monomers. After copolymerization, obtained copolyesters exhibited a wide range of thermal and mechanical properties. They could be complete amorphous plastics or high performance elastomers.

33, 39

This indicates the potential of FDCA as a bio-based

building block for polymer synthesis.

Figure 1. Dependence of melting temperature and tensile strength on the carbon number of diols for FDCA-based polyesters As we know, not only the skeletal structure of diols, but also the substitutions on them significantly influence polyesters’ properties. For instance, Ahn and his coworkers synthesized some copolyesters with the same skeletal structure as poly(ethylene glycol-co-1,4-cyclohexanedimethanol terephthalate) (PETG) from 1,2propanediol and 2,3-butanediol. They found that the number of lateral methyl groups showed remarkable influence on Tg and mechanical properties of synthesized copolyesters.41 Papageorgiou and his coworkers investigated properties of poly(methylpropylene furanoate) (PMePF) and reported significantly different thermal properties from those of PPF.

42

In addition, it is well known that free volume, free volume

distribution and chain motions are main parameters determining the permeability of a barrier polymer. 43 As for the FDCA polyesters, their good gas barrier properties have 4

ACS Paragon Plus Environment

Page 4 of 34

Page 5 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

been attributed to the frustrated furan ring-flipping and ring polarity, 44 and the lateral groups on the main chain of FDCA polyesters will undoubtedly influence their free volume as well as chain motions, and then the barrier properties. However, to our knowledge, the effects of pendant groups on the properties of FDCA polyesters has not been investigated. In this work, the aliphatic diols with same skeletal structures, but different pendant groups, such as 1,3-propanediol, 2-methyl-1,3-propanediol, neopentyl glycol, 2methyl-2-propyl-1,3-propanediol

and

2-ethyl-2-butyl-1,3-propanediol

were

polymerized with FDCA. Properties of obtained polyesters in terms of crystallization behaviors, dynamic mechanical properties, gas barrier properties, segment motions, and free volume were investigated by DSC, TGA, DMA, gas permeability tester, rheological analyzer and positron annihilation lifetime spectroscopy (PALS). As shown in Scheme 1, based on the regular change of substituent groups on 1,3-propanediol skeleton, the effects of various 1, 3-propanediols were studied and then a new window for the modification of FDCA polyesters would be opened, which is the main objective of this work.

Scheme 1 Illustration for the synthesis of polyesters from FDCA and different diols 5

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

EXPERIMENTAL SECTION Materials. Antimony trioxide (III, 99.99%), zinc acetate (99%) and trifluoroacetic acid were all obtained from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China). 2,5furandicarboxlic acid (2,5-FDCA) was purchased from Ningbo Jisu New Material Technology Co., Ltd. (Ningbo, China). 1,3-propanediol (PDO) (99%), 2-methyl-1,3propanediol (MPD) (99%), neopentyl glycol (NPG) (99%), 2-methyl-2-propyl-1,3propanediol (99%) and 2-ethyl-2-butyl-1,3-propanediol (98%) were all purchased from Aladdin Reagent Co. Ltd (Shanghai, China). All the chemicals were used as received without further purification.

Synthesis of different polyesters Polyesters were synthesized through the conventional two-step polycondensation procedure, involving polymerization of FDCA with PDO, MPD, NPG, 2-methyl-2propyl-1,3-propanediol as well as 2-ethyl-2-butyl-1,3-propanediol. Similar to the previously reported procedure,2,

4

FDCA was converted into dimethyl furan-2, 5-

dicarboxylate (DMFD) at first and the molar ratio of diols to DMFD was fixed at 1.6. Into a 1000 mL three-necked round bottom flask equipped with a mechanical stirrer, the predetermined diols (0.48mol), DMFD (0.30mol) and the first portion of catalyst (Zinc acetate, 0.2 mol% based on DMFD) were added. Before heating, a vacuum of 0.1 kPa was applied on the mixture and purged with N2 gas for three times to remove air. Then the reaction was conducted at 180-200 oC for 3-4h under a constant nitrogen purge and subsequently added the second portion of catalyst (antimony trioxide (III), 0.15 mol% based on DMFD). After that, the temperature was increased up to 230-240 oC and the reaction pressure was gradually reduced to 10-30 Pa. In the following 2.5-3.0 h, the torque value of the stirrer was measured and reached the desired value. Finally, the reaction pressure was adjusted to the normal atmospheric pressure by the slow introduction of N2. The target polyesters (PPF, PMF, PNF, PMPF and PEBF) were obtained and the reaction conditions are listed in Table 1 below.

6

ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Table 1 Feeding ratio, reaction conditions and molecular weight for PPF, PMF, PNF, PMPF and PEBF

Sample

DMFD (mol)

Esterification

Polycondensation

GPC

Diol

T

Time

T

Time

Mn

Mw

(mol)

(oC)

(h)

(oC)

(h)

(g/mol)

(g/ mol)

Ð

PPF

0.30

0.48

180

3.0

230

3.0

27400

43400

1.6

PMF

0.30

0.48

180

3.0

230

3.0

30200

53800

1.8

PNF

0.30

0.48

180

3.0

230

3.0

45900

79800

1.7

PMPF

0.30

0.48

200

4.0

240

2.5

19500

37700

1.9

PEBF

0.30

0.48

200

4.0

240

2.5

26400

44900

1.7

Characterization The molecular weight and molecular weight distribution of synthesized polyesters were measured at 40 oC using GPC (Agilent PL-GPC220) equipped with two columns (PLgel 5 μm Mixed-D 300*7.5 mm). Chloroform and 2-chlorophenol mixture with the volume ratio of 9/1 was used as the solvent. The concentration of polyesters was about 1 mg/mL. Chloroform was used as the mobile phase and the flow rate was 1 mL/min. The equipment was calibrated with polystyrene standards with the molecular weight range of 3070 to 258000 g/mol. The chemical structures of obtained polyesters were determined by NMR using a Bruker AVIII400 NMR spectrometer. CF3COOD was used as the solvent and the measurement was conducted at 25 oC. Tetramethylsilane (TMS) was used as the internal standard. DSC measurements were conducted on a Mettler-Toledo DSC I differential scanning calorimeter under nitrogen atmosphere. The sample was first heated to 250 oC with a heating rate of 10 oC/min and maintained at this temperature for 3 min. Then it was cooled down to 25 oC at a rate of 10 oC /min and heated again to 250 oC with the same heating rate. All heating and cooling scan curves were recorded for analysis. TGA (Mettler-Toledo TGA/DSC thermogravimetric analysis) was taken to determine thermal stabilities of synthesized polyesters. About 6-8 mg sample was placed in a 7

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ceramic furnace and heated from 50 to 800 oC at a rate of 20 oC/min. The measurement was conducted under nitrogen with a flow rate of 50 mL/min. Dynamic rheological measurement was conducted on an Anton Paar Physica MCR301. Diameters of upper and lower plates were 25 and 50 mm, respectively. The measurement was performed at 5 °C/min in a shear mode with the frequency of 1 Hz and strain of 0.1%. Dynamic mechanical analysis (DMA) was conducted on a METTLER-TOLEDO dynamic mechanical analyzer (DMA/Q800) at a fixed frequency of 1 Hz. The samples with a dimension of 8.0× 6.0×1.0 mm were prepared by hot pressmolding. The sample was scanned from -100 to 150 oC with a heating rate of 3 oC/min. Barrier properties were investigated by Labthink VAC-V2 gas permeability tester. High purity O2 and CO2 (99.99%) were used. Amorphous polyesters round films with diameter of 97 mm were prepared by melt-press/quench procedures. The testing area was 38.5 cm2 and the test was conducted in the range of 0.05 to 50000 cm3/m2·24h·0.1MPa with the relative humidity (RH) of 50% at 30 oC. In order to ensure accuracy, specimens were tested three times and averaged values are reported. Positron Annihilation Lifetime Spectroscopy (PALS) was performed on a typical two detector fast-slow coincidence timing spectrometer. The measurement throughput was 2×106 events per spectrum and the time resolution was 0.23 ns. 2 sheets with the dimension of 1.5cm×1.5cm×1 mm were stacked on either side of the source to ensure the Ps annihilated in the polyester film. A Gaussian resolution function with exponential flanks was convolved with a discrete lifetime exponential spectrum, and the entire lifetime histogram of positron and positronium decays was fitted.

8

ACS Paragon Plus Environment

Page 8 of 34

Page 9 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

RESULTS AND DISCUSSION Synthesis and Structure Confirmation of Different Polyesters As shown in Scheme 1, different FDCA-based polyesters, PPF, PMF, PNF, PMPF and PEBF were synthesized through a two-step procedure consisting of transesterification and polycondensation. In order to restrain coloration, instead of esterification reaction with diols, FDCA was converted into dimethyl furan-2,5-dicarboxylate (DMFD) at first, 2, 4

and then the transesterification and polycondensation reactions were conducted.

Considering the different boiling temperatures and reaction activities of varied diols, slightly different reaction conditions were applied for different polyesters (Table 1). The obtained polyesters’ weight-average molecular weight (Mw) varied from 37700 to 79800 g/mol, with the molar-mass dispersity of 1.6-1.9 based on the GPC measurement shown in Table 1. The molecular weights of PMPF and PEBF were difficult to be further increased by prolonging the polycondensation time because of the relatively high boiling point of 2-methyl-2-propyl-1,3-propanediol (5.2kPa, 230 oC) and

2-

ethyl-2-butyl-1,3-propanediol (0.67 kPa, 160 oC), which were difficult to be distilled out under the polycondensation condition in this work. Figure 2 shows the 1H-NMR spectra and characteristic peak assignments of PPF, PMF, PNF, PMPF and PEBF. The chemical shift of CH in furan ring (f) is shown at 7.13-7.20 ppm and the CH2 in different diols connected with ester bonds show characteristic peaks in the range of 4.19-4.45 ppm. The peak for the CH2 (b) in 1, 3propanediol, which is not connected to the ester bond, appears at 2.15 ppm. The peaks shown at 1.0 ppm (d) are assigned to the pendant CH3 in PMF, PNF and PMPF. As for PMPF, the characteristic peaks at 1.32, 1.24 and 0.76 ppm are assigned to the propyl unit (g, e, h). The signals at 1.30 (i) and 0.68 ppm (j) are associated with the pendant ethyl and the peaks at 1.39 (m), 1.37 (l), 1.30 (k) and 0.75 ppm (n) are for the side butyl in PEBF.

9

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2 1H-NMR spectra and characteristic peak assignments of synthesized polyesters

The 13C-NMR spectra of PPF, PMF, PNF, PMPF and PEBF are shown in Figure 3. The chemical shifts of aromatic carbons in FDCA units appeared at 119 (peak e) and 146 ppm (peak f). The peak at 160 ppm, which overlapped with the solvent, is assigned to the carbonyl carbon connected with furan ring. As for the signals shown in the range of 5-40 ppm, they are all assigned to the aliphatic carbons in different diols units. Based on NMR results, chemical structures of synthesized polyesters were confirmed.

10

ACS Paragon Plus Environment

Page 10 of 34

Page 11 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 3 13C-NMR spectra and characteristic peak assignments of synthesized polyesters

Thermal Properties of Synthesized Polyesters Thermal properties of synthesized polyesters were probed by DSC and the first heating scans, cooling scans and second heating scans from 25 to 250 oC at a rate of 10 oC/min are shown in Figure 4. The values of Tg, Tm, melting enthalpy (ΔHm),crystallization temperature (Tc), cold crystallization temperature (Tcc) and enthalpy (ΔHc) are summarized in Table 2. In Figure 4(a), melting and crystallization peaks are not observed during the first heating scans for PPF, PMF, PMPF and PEBF. But PNF showed a melting peak at 202.9 oC (Tm’) with ΔHm’ of 33.0 J/g, and crystallization temperature (Tcc) at 145.3 oC with ΔHcc of 32.6 J/g. Figure 4(b) shows the cooling scans from 250 oC to 25 oC for all the samples and no crystallization peak is observed. In Figure 4c, PPF shows Tm at 173.3 oC with ΔHm of 0.9 J/g, indicating its weak crystallization. In Lotti and Siracusa’s work,10, 24 they also reported the DSC results of PPF. During the second scan, neither 11

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PPF powder nor film showed obvious Tm, which also indicated the weak crystallizability of PPF. When H atom on the middle CH2 in 1,3-propanediol was substituted by methyl (1,3-propanediol was replaced by 2-methyl-1,3-propanediol), PMF became an amorphous polyester indicated by the absent melting peak in the heating scan. However, when the other H was also replaced by a methyl group and the diol was changed into NPG, the obtained PNF unexpectedly showed obvious Tm at 202.9 oC with ΔHm of 33.0 J/g. As we know, the value of ΔHm can be taken as an indicator to compare the crystallizability of semi-crystalline polyesters. Usually, higher ΔHm means better crystallizability under the same condition. For example, the weak crystalline PEF and PPF shows ΔHm of only 0.5-0.9 J/g. 4 While for PBF, known as the FDCA polyester with satisfied crystallizability, it demonstrates high ΔHm up to 32 J/g. 4, 45 In

this work, the ΔHm of 33.0 J/g for PNF also indicated the high crystallizability of

PNF. The reason might be that NPG was more symmetric than MPD, which efficiently improved the chains’ regularity and then improved the capacity of crystallization. When compared with PPF, the higher crystallizability of PNF was once ascribed to the presence of short polymer chains, indicated by the large polydispersity index (Ð) of 4.0 in previous work. 24 However, in this work, the Ð of all the synthesized polyesters was in the range of 1.6-1.9 (Table 1). Therefore, the reason for the good crystallizability of PNF is worthy of further investigation. As for PMPF and PEBF, when the diols lost the structural symmetry again, they performed as a complete amorphous polyester accordingly. In addition, as shown in Figure 4(c), from PPF to PNF, and then to PEBF, their Tg showed a decreasing tendency after the first increase. Tg of PNF was high up to 70.3 oC, 15 oC higher than that of PPF. The reason was ascribed to the presence of side methyl group in place of hydrogen atoms in PNF, which reduced the chain flexibility. 24

When the length of pendant substitution was long enough, it would exert an internal

plasticizing effect and then lead to a decreased Tg. 46, 47

12

ACS Paragon Plus Environment

Page 12 of 34

Page 13 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 4 DSC curves of synthesized polyesters. (a) First heating scan from 25 to 250 oC;

(b) Cooling curves from 250 to 25 oC; (c) Second heating scans from 25 to 250 oC.

The heating or cooling rate is 10 oC /min; (d) TGA curves of PPF samples in nitrogen.

The thermal stability of above polyesters was investigated by TGA (Figure 4d). The temperature at which the weight loss reached 5% of its initial weight (T5%) and the temperature at which the maximum decomposition rates occurred (Td,max) are summarized in Table 2. All the samples demonstrated good thermal resistance regardless of the change in pendant groups, indicated by T5% and Td,max (higher than 370 and 410 oC, respectively). It was easy to notice that the substituent groups on the 1,3propanediol unit showed obvious impact on the thermal stability of synthesized polyesters. When the substituent group was changed from H to CH3 and then propyl, T5% increased from 381 oC for PPF to 412 oC for PMPF. Furthermore, Td,max also showed a 48 oC increment from 413 oC for PPF to 461 oC for PMPF. According to literatures, T5% of petroleum-based PET and PPT are in the range of 370 to 403 oC, while Td,max of PET and PBT ranges from 400 to 440 oC. 48, 49 In this work, the T5% and Td,max for PPFs 13

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 34

were in the similar temperature region, indicating their excellent thermal stability. 50, 51 In addition, the introduction of lateral groups obviously improved the thermal stability of PPFs. As previously reported, 50-53 the thermal degradation of polyesters was highly dependent on the nature of diols, and the widely accepted mechanism for the decomposition of diol subunit was the β-scission. In this work, it might be that the lateral groups depressed the β-scission reaction, leading to better thermal stability. Table 2 Thermal properties and thermal stabilities of PPFs DSC

TGA

First heating scan

Second heating scan

N2

Sample Tcc

ΔHcc

Tm’

ΔHm’

Tg

Tcc

ΔHcc

Tm

ΔHm

Td,5%

Td,max

(oC)

(J/g)

(oC)

(J/g)

(oC)

(oC)

(J/g)

(oC)

(J/g)

(oC)

(oC)

PPF

nd

nd

nd

nd

55

nd

nd

173

0.9

381

413

PMF

nd

nd

nd

nd

54

nd

nd

nd

nd

390

426

PNF

145

33

203

33

70

150

31

201

32

411

452

PMPF

nd

nd

nd

nd

54

nd

nd

nd

nd

412

461

PEBF

nd

nd

nd

nd

45

nd

nd

nd

nd

409

451

Barrier Properties of Synthesized Polyesters The amorphous, unoriented polyester films were prepared by the melt-press/quench procedure for barrier properties evaluation. During this process, PMPF and PEBF were too brittle to be processed into films. Therefore, only the barrier properties of PPF, PMF and PNF, which showed similar tensile properties (Supporting Information Figure S4 and Table S1), were investigated by Labthink VAC-V2 gas permeability tester. The detailed CO2 and O2 permeability coefficient values are shown in Table 3. For easy expression, the Barrier Improvement Factor (BIFP) widely used for gas permeability comparison, 28, 29, 54 which represents the permeability coefficient of CO2 or O2 in PET divided by the CO2 or O2 permeability in other polymers, was employed to quantitatively compare the barrier properties of PPF, PMF and PNF. As shown in Table 3, the CO2 BIFp of PPF, PMF and PNF were 8.1, 7.5 and 4.2, while for O2 BIFp were 14

ACS Paragon Plus Environment

Page 15 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

6.7, 6.0 and 4.0, respectively. These results indicated that from PPF to PMF, then to PNF, the gas-barrier properties decreased gradually. In our previous work,

2

the gas

barrier property of PEF was investigated under the same condition. Compared to PEF, it was found that the barrier of PPF to CO2 was poorer. However, PPF and PMF showed better barrier against O2 with respect to PEF, indicated by the O2 BIFp of 6.7 for PPF and 6.0 for PMF vs. 5.5 for PEF.

2

Overall, PPF, PMF and PNF all showed great

potential to be used as high barrier materials. It has been widely accepted that the free volume, free volume distribution, and chain dynamics are the main parameters determining the permeability of a barrier polymer. As for the FDCA-based polyesters, it has been accepted that PEF possesses higher free volume compared to PET due to the non-axisymmetric furan ring and Burgess et al suggested the suppressed furan-ring flip as the main reason explaining their high gas barrier properties. 43 In this work, PPF, PMF and PNF possessed the same FDCA unit and the contribution of FDCA units should be the same. Therefore, the different pendent groups on diols should be responsible for their difference in CO2 and O2 BIFp values. In the following section, the molecular segment motions and free volumes of PPF, PMF and PNF are investigated by dynamic mechanical analysis and Positron Annihilation Lifetime Spectroscopy.

Table 3 CO2 and O2 permeability coefficient of PPF, PMF and PNF Temperature

CO2 permeability

(oC)

coefficient (barrer b)

PETc

30

0.13

1

0.060

1

PPF

30

0.016

8.1

0.009

6.7

PMF

30

0.017

7.5

0.010

6.0

PNF

30

0.031

4.2

0.015

4.0

PEF 2

30

0.010

13

0.011

5.5

Samplea

a

O2 permeability coefficient (barrer b)

The test performed at low pressure (≤0.1001 MPa);

b 1 barrer=10-10 c

CO2 BIFP

cm3·cm/cm2·s·cmHg.

PET was synthesized in our laboratory with the intrinsic viscosity of 0.82 15

ACS Paragon Plus Environment

O2 BIFP

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Dynamic Mechanical Properties Dynamic mechanical properties of PPF, PMF and PNF were investigated by DMA and the dependence of storage modulus (E’) and tan δ (ratio of the loss modulus to storage modulus) on temperature are shown in Figure 5. Figure 5(a) shows the E’ as a function of temperature from -90 to 130 oC for all the polyesters and two obvious decrement stages were observed. The first stage occurred in the range of -90 to 50 oC, which should be associated with the short-range segmental motion and reported as the β relaxation. The second stage was observed from 50 to 130 oC, where the Tg was recognized by the rapid drop in E’ (α relaxation). In Figure 5(b), the Tg of PPF was at 72.9 oC, while it decreased to 63.2 oC when 1,3-propanediol was replaced by 2-methyl-1,3-propanediol, and the Tg of PNF increased up to 88.5 oC. These results were in line with the Tg values determined by DSC, although there were some deviations for the same specimens tested by DSC or DMA. Based on the gas solution-diffusion theory,

55

permeability (P) is a product of

solubility coefficient (S) and diffusivity (D) (P=D×S). In this work, β relaxation is our primary interest, which is typically related with the small-scale molecular motions and then determined by the diffusivity of polymers.56 In Figure 5(b), the β relaxation of PPF, PMF, and PNF was enlarged. It was noted that the peak temperature of β relaxation decreased from -14.4 oC for PMF to -25.5 oC for PNF, and the β peak areas increased accordingly. Light and his coworker 57 reported a direct correlation between segmental motions related with β peak area and gas diffusivity. They observed that the area under β peak and a peak temperature of β relaxation would increase or decrease with the addition of mobility-enhancing or mobility-suppressing unit. Burgess et al compared the β relaxation of PET and PEF. They also attributed the higher β peak temperature and suppressed magnitude to the more restricted motions of furan ring and carbonyl, which is one of the reasons explaining the better barrier properties of PEF. 43 Recently, Soccio and Ezquerra investigated the chemical structure-subglass relaxation dynamics relationship for PNF and poly(trimethylene 1,4-cyclohexanedicarboxylate) via dielectric spectroscopy. They also believed that the secondary relaxation and the lower molecular mobility of furan-based polymers were connected with the low gas 16

ACS Paragon Plus Environment

Page 16 of 34

Page 17 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

permeability of these materials.58 As discussed in the above section, from PPF to PNF, their gas barrier properties followed the trend of PPF>PMF>PNF. The increased β relaxation magnitude associated with the substituent groups on 1,3-propanediol skeleton might be one of the reason.

Figure 5 Storage modulus (a) and tan δ (b) as a function of temperature for PPF, PMF and PNF

17

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 34

Rheological analysis (RA) for PPF, PMF, and PNF Melt rheological analysis is a good method to investigate the structural influence on chain dynamics of polymers. Figure 6 (a) shows the representative plots of storage (Gʹ) and loss (Gʺ) modulus as a function of frequency at the reference temperature of 270 °C. As shown in Figure 6 (a), Gʺ was higher than Gʹ in the range of 10-1-103 rad/s, which meant that the melts were in the viscous flow state. In the frequency of 10-1000 rad/s, PPF, PMF and PNF showed similar Gʹ and Gʺ, and they linearly increased with frequency. The slope was determined to be 0.95-0.97 for Gʺ and 1.39-1.42 for Gʹ in the frequency range of 10-100 rad/s. For the typical linear polymer melt, they usually show the Gʺ slope of ~1.0 and the Gʹ slope of ~2.0, which is well consistent with our results. 59

This result indicates that the addition of –CH3 pendant groups didn’t change the flow

behaviors of PMF and PNF from linear polyester to branched polyester. In the lower frequency range of 1-10 rad/s, PPF and PMF showed similar Gʹ and Gʺ. But for PNF, its Gʹ and Gʺ were obviously higher than those of PPF and PMF. The Gʹ and Gʺ were influenced by the combined action of molecular entanglement, disentanglement and orientation. The Gʺ, which is related to the viscous response, refers to the energy dissipated by the internal friction of molecule chain mobility and Gʹ is related to the elastic response, which is attributed to the rigidity of the molecule chain.

60

The

relatively higher Gʹ and Gʺ of PNF at the same frequency should be assigned to the increased bulky effect and the more energy dissipation at viscous flow state caused by the more methyl pendant groups.

18

ACS Paragon Plus Environment

Page 19 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 6 (a) storage modulus (G') and loss modulus (G''); (b) complex viscosity versus angular frequency of PPF, PMF and PNF (Tr = 270 °C)

As we know, the viscosity determined by rheological analysis is the complex viscosity ƞ*. And the following equations 1-3 describe the relationship between ƞ*, ƞʹ, 19

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 34

ƞ'', Gʹ and Gʺ. The ƞʹ is called dynamic complex viscosity and is related to the loss modulus (G''), while ƞ'' is related to the storage modulus (G').

ƞ * = ƞʹ - iƞʺ (1) ƞʹ = Gʹ/ω (2) ƞʺ = Gʺ/ω (3)

When the melt becomes into Newtonian fluid, the viscosity will not change with the shear rate and the equilibrium state viscosity is the so-called zero shear viscosity.

61

Figure 6(b) shows the complex viscosity curves of PPF, PMF and PNF. In the frequency range of 1-1000 rad/s, the viscosities of PPF, PMF and PNF did not change with the frequency and the zero-shear viscosities of them were all about 20 Pa·s. The reason was that, different from the long-chain lateral groups, the molecular chain entanglement and dissociation caused by the pendant –CH3 was not obvious. In addition, the molecular weight of PNF and PMF was a little higher than that of PPF, which made the viscous flows of PNF and PMF a little more difficult.

Figure 7 The shift factor data of PPF, PMF, and PNF in 230- 270 °C (Tr = 270 °C)

20

ACS Paragon Plus Environment

Page 21 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

PPF, PMF and PNF should obey the Williams-Landel-Ferry (WLF) equation. The shift factor data (aT) (Equation 4), evaluated from the master curve by shifting the values along the frequency (time) scale, was obtained by the Time-Temperature equivalence principle using the TRIO software. The shift factor aT curves for PPF, PMF and PNF are shown in Figure 7. Using Williams-Landel-Ferry (WLF) equation to fit the aT curves, the linearly dependent coefficient of 0.9899 well verified our data accuracy. The WLF parameters C1 and C2 could be obtained directly from the TRIO software and are dependent on Tref. The WLF parameters Cg1 and Cg2 as well as the free volume fraction (FFV) at Tg (fg) can be calculated by the following equations (57):

― 𝐶1(𝑇 ― 𝑇𝑟𝑒𝑓)

log 𝑎𝑇 = 𝐶2 + (𝑇 ― 𝑇𝑟𝑒𝑓) 𝐶1𝐶2

Cg1 = 𝐶2 + (𝑇g ― 𝑇ref) Cg2 = 𝐶2 + (𝑇g ― 𝑇ref) 𝐵

(4) (5) (6) (7)

𝑓g = 2.303𝐶g

1

The fg and related parameters are shown in Table 4. It was noted that FFV at Tg gradually increased from 0.0044 for PPF to 0.0103 for PNF. That should be attributed to the bulky effect of the pendant group leading to higher free volume. From PPF to PNF, H atoms in HOCH2CH2CH2OH were gradually replaced by –CH3, and with the increasing mobile terminals, the bulky effect appeared and then led to the larger FFV. This result was consistent with the β relaxation observed by the above DMA analysis. In Finkelshtein’s work,

62

they also reported the larger space occupied by the (CH3)3

group compared with that of -CH3 group. Generally, the higher FFV is correlated to the higher gas permeability in a polymer.44 In this work, from PPF to PNF, their gas barrier properties followed the trend of PPF>PMF>PNF, which was in line with the FFV sequence.

21

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 34

Table 4 WLF Parameters and Fractional Free Volumes of PPF, PMF and PNF o

o

C1

C2 (K)

Tg( C)

Tref( C)

Cg1

Cg2 (K)

fg

PPF

4.52

225

55.4

270

97.8

10.4

0.0044

PMF

4.25

230

54.4

270

67.9

14.4

0.0064

PNF

4.58

248

70.3

270

42.2

24.3

0.0103

Positron Annihilation Lifetime Spectroscopy (PALS) It is well known that positron can exist as free positron or positronium (Ps) in polymers and the lifetime of positronium is determined by the free volume cavity size and spin state. The longest lifetime component τ3 (in the range 2-3 ns), characteristic for the ortho-Positronium (o-Ps), is an indicative of free volume in an amorphous material. Therefore, PALS is a useful technique to probe the microstructure of polymers. Figure 8 (a) and (b) show the average o-Ps lifetime, τ3 and the corresponding o-Ps intensity I3. The free volume size can be obtained by the relationship of τ3 and radium (R) by the following equation 8. 63

[

1

𝑅

𝜏3 = 2 1 ― 𝑅 + 𝛥𝑅 +

(2𝜋1 )sin (𝑅2𝜋𝑅 + 𝛥𝑅)]

―1

(8)

where ΔR=0.1656 nm is the fitted electron layer thickness taken from literature. 64 The intensity, I3, is proportional to the number of holes. As shown in Figure 8 (a), from PPF to PNF, τ3 increased with the number of lateral –CH3 groups. Additionally, the values of R also increased as well. In Figure 8(b), I3 also increased from 14.67% for PPF to 20.74% for PMF and then 23.22% for PNF. The increased intensity indicated the increased amounts of holes. R and I were related to the fraction free volume (FFV) of materials, as in equation 9: 22

ACS Paragon Plus Environment

Page 23 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

𝐹𝐹𝑉 = 𝐶𝑉𝐹𝐼3

(9)

C is an arbitrarily scaling factor for a spherical cavity (C = 1.5), 𝑉𝐹 =

4𝜋𝑅3 3

is the

average free volume with the assumption of fractional free volume is around 10%.

65

Figure 8 (c) shows the calculated FFV values. From PPF to PNF, FFV increased from 0.0142 for PPF to 0.0274 for PNF. This sequence was in good agreement with above rheological analysis result. Based on these results, it can be concluded that, from PPF to PNF, the decreased gas barrier properties should be related to the increased β relaxation and FFV, after more lateral -CH3 group was introduced.

23

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8 (a) o-Ps lifetime and radium of the size of the free volume holes in PPF, PMF and PNF; (b) o-Ps intensity, indicative of the population of the holes in PPF, PMF and PNF; (c) Fractional free volume of PPF, PMF and PNF. (Dotted lines were plotted only to aid eye)

CONCLUSIONS The influence of substituent groups on propanediol on the performance of PPF was investigated via synthesis of different polyesters with same skeletal structure. From PPF to PNF and then to PEBF, their Tg showed a decreasing tendency after the first increase. PPF exhibited Tm at 173.3 oC with ΔHm of 0.9 J/g and PMF was a completely amorphous polyester. However, PNF unexpectedly showed ΔHm high up to 33.0 J/g, indicating its good crystallizability. These results revealed that not only the amount of the substituent groups, but also the structural symmetry of diols played significant roles in determining the thermal properties of resulted polyesters. As for the gas barrier properties, they followed the trend of PPF>PMF>PNF and the reason was attributed to the increased β relaxation and FFV after more lateral -CH3 groups were introduced. Based on these results, it can be concluded that, not only the amount of substituent groups but also the structural symmetry of diols significantly influenced the properties of polyesters. Through the manipulation of lateral groups on diols, the FDCA-based polyesters with varied properties could be achieved. 24

ACS Paragon Plus Environment

Page 24 of 34

Page 25 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

ASSOCIATED CONTENT Supporting Information NMR spectra of dimethyl furan-2, 5-dicarboxylate; GPC curves for synthesized polyesters; Tensile properties of PPF, PMF and PNF

AUTHOR INFORMATION Corresponding Author E-mail: [email protected]. [email protected] ORCID Xiaoqing Liu: 0000-0002-7417-1326

Notes The authors declare no competing financial interest.

ACKNOWLEDGEMENTS The authors greatly thank the financial support from National Key Research and Development Program of China (2018YFD0400700), Chinese Academy of Sciences (Key Research Program ZDRW-CN-2016-1), Natural Science Foundation of Shandong Province of China (ZR2016BL25), Medical Health Brand Discipline Project of Ningbo (PPXK2018-02), Sinopec Technology Support Project (2017002-4) and Project of Ningbo Natural Science Foundation (No. 2018A610032).

REFERENCES (1) Jong, E.; Dam, M. A.; Sipos, L.; Gruter, G. J. Furandicarboxylic acid (FDCA), a versatile building block for a very interesting class of polyesters. ACS Symposium 2012, 1105, 1-13, DOI 10.1021/bk-2012-1105.ch001. (2) Wang, J. G.; Liu, X. Q.; Zhang, Y. J.; Liu, F.; Zhu, J. Modification of poly(ethylene 2,5-furandicarboxylate) with 1,4-cyclohexanedimethylene: influence of composition 25

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

on

mechanical

and

barrier

properties.

Polymer

Page 26 of 34

2016,

103,

1-8,

DOI

10.1016/j.polymer.2016.09.030 (3) Knoop, R. J.; Vogelzang, W.; Haveren, J. V.; Es, D. S. High molecular weight poly(ethylene-2,5-furanoate); critical aspects in synthesis and mechanical property determination. J. Polym. Sci. Part A: Polym. Chem. 2013, 21, 4191-4199, DOI 10.1002/pola.26833. (4) Wang, J. G.; Liu, X. Q.; Jia, Z.; Liu, Y.; Sun, L. Y.; Zhu, J. 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, 305-320, DOI 10.3390/polym9090305. (5) Berkel, J. G; Guigo, N.; Kolstad, J. J.; Sipos, L.; Wang, B.; Dam, M. A.; Sbirrazzuoli, N. Isothermal crystallization kinetics of poly(ethylene 2,5-furandicarboxylate). Macromol. Mater. Eng. 2015, 300, 466-474, DOI 10.1002/mame.201400376. (6) Martino, L.; Guigo, N.; Berkel, J. G.; Kolstad, J. J.; Sbirrazzuoli, N. Nucleation and self-nucleation of bio-based poly(ethylene 2,5-furandicarboxylate) probed by fast scanning calorimetry. Macromol. Mater. Eng. 2016, 301, 586-596, DOI 10.1002/mame.201500418. (7) Wang, J. G.; Liu, X. Q.; Jia, Z.; Liu, Y.; Sun, L. Y.; 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, DOI 10.1002/pola.28706. (8) 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. Part A: Polym. Chem. 2012, 50, 1026-1036, DOI10.1002/pola.25859. (9) Vannini, M.; Marchese, P.; Celli, A.; Lorenzetti, C. Fully bio-based poly(propylene 2,5-furandicarboxylate) for packaging applications: excellent barrier properties as a function of crystallinity. Green Chem. 2015, 17, 4162-4166, DOI 10.1039/c5gc00991j. (10) Guidotti, G.; Soccio, M.; Lotti, N.; Gazzano, M.; Siracusa, V.; Munari, A. poly(propylene

2,5-thiophenedicarboxylate)

vs.

poly(propylene

2,5-

furandicarboxylate): two examples of high gas barrier bio-based polyesters. Polymers 2018, 10, 785, DOI 10.3390/polym10070785. 26

ACS Paragon Plus Environment

Page 27 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(11)Tsanaktsis, V.; Terzopoulou, Z.; Nerantzak, M.; Papageorgiou, G. Z.; Bikiaris, D. N. New poly(pentylene furanoate) and poly(heptylene furanoate) sustainable polyesters from diols with odd methylene groups. Mater. Lett. 2016, 178, 64-67, DOI 10.1016/j.matlet.2016.04.183. (12) 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, DOI 10.1016/j.eurpolymj.2016.08.004. (13) Zhu, J.; Cai, J.; Xie, W.; Chen, P.; Gazzano, M.; Scandola, M.; Gross, R. A. Poly(butylene 2,5-furan dicarboxylate), a biobased alternative to PBT: synthesis, physical properties, and crystal structure. Macromolecules 2013, 46, 796-804, DOI 10.1021/ma3023298. (14) Papageorgiou, G. Z.; Tsanaktsis, V.; Papageorgiou, D. G.; Exarhopoulos, S.; Papageorgiou, M.; Bikiaris, D. N. Evaluation of polyesters from renewable resources as alternatives to the current fossil-based polymers. Phase transitions of poly(butylene 2,5-furan-dicarboxylate).

Polymer

2014,

55,

3846-3858,

DOI

10.1016/j.polymer.2014.06.025. (15) Cai, X.; Yang, X.; Zhang, H.; Wang, G. Aliphatic-aromatic poly(carbonate-coester)s containing biobased furan monomer: synthesis and thermo-mechanical properties. Polymer 2018, 134, 63-70, DOI 10.1016/j.polymer.2017.11.058. (16) Soccio, M.; Martínez-Tong, D. E.; Alegría, A.; Munari, A.; Lotti, N. Molecular dynamics of fully biobased poly(butylene 2,5-furanoate) as revealed by broadband dielectric

spectroscopy.

Polymer

2017,

128,

24-30,

DOI

10.1016/j.polymer.2017.09.007. (17) Tsanaktsis, V.; Papageorgiou, G. Z.; Bikiaris, D. N. A facile method to synthesize high-molecular-weight biobased polyesters from 2,5-furandicarboxylic acid and longchain diols. J. Polym. Sci. Part A: Polym. Chem. 2015, 53, 2617-2632, DOI 10.1002/pola.27730. (18) Zhang,D.; Dumont, M. Advances in polymer precursors and bio-based polymers synthesized from 5-hydroxymethylfurfural. J. Polym. Sci. Part A: Polym. Chem. 2017, 27

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 34

55, 1478-1492, DOI 10.1002/pola.28527. (19) Papageorgiou, G. Z.; Guigo, N.; Tsanaktsis, V.; Papageorgiou, D. G.; Exarhopoulos, S.; Sbirrazzuoli, N.; Bikiaris, N. D. On the bio-based furanic polyesters: synthesis and thermal behavior study of poly(octylene furanoate) using fast and temperature modulated scanning calorimetry. Eur. Poly. J. 2015, 68, 115-127, DOI 10.1016/j.eurpolymj.2015.04.011. (20) 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-furandicarboxylate).

Eur.

Polym.

J.

2015,

67,

383-396,

DOI

10.1016/j.eurpolymj.2014.08.031. (21) Tsanaktsis, V.; Bikiaris, D. N.; Guigo, N.; Exarhopoulos, S.; Papageorgiou, D. G.; Sbirrazzuoli, N.; Papageorgiou, G. Z. Synthesis, properties and thermal behavior of poly(decylene-2,5-furanoate): a biobased polyester from 2,5-furan dicarboxylic acid. RSC Adv. 2015, 5, 74592-74604, DOI 10.1039/c5ra13324f. (22) Shen, Y.; Yao, B. B.; Yu, G.; Fu, Y.; Liu, F. S.; Li, Z. B. Facile preparation of biobased polyesters from furandicarboxylic acid and long chain diols via asymmetric monomer strategy. Green Chem. 2017, 19, 4930-4938, DOI 10.1039.C7GC02081C. (23) Papageorgiou, D. G.; Guigo, N.; Tsanaktsis, V.; Exarhopoulos, S.; Bikiaris, D. N.; Sbirrazzuoli, N.; Papageorgiou. G. Z. Fast crystallization and melting behaviour of a long spaced aliphatic furandicarboxylate biobased polyester, poly(dodecylene 2,5furanoate). Ind. Eng. Chem. Res. 2016, 55, 5315-5326, DOI 10.1021/acs.iecr.6b00811. (24) 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(9), 1028, DOI 10.3390/ma10091028. (25)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, DOI 28

ACS Paragon Plus Environment

Page 29 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

10.1039/c5py01367d. (26) Papageorgiou, G. Z.; Tsanaktsis, V.; Bikiaris, D. N. Synthesis of poly(ethylene furandicarboxylate) polyester using monomers derived from renewable resources: thermal behavior comparison with PET and PEN. Phys. Chem. Chem. Phys. 2014, 16, 7946-7958, DOI 10.1039/c4cp00518j. (27) Runt, J.; Miley, D. M.; Zhang, X.; Gallagher, K. P.; Featers, K. M.; Fishburn, J. Crystallization of poly(buty1ene terephthalate) and its blends with polyarylate. Macromolecules 1992, 25, 1929-1934, DOI 10.1021/ma00033a015. (28) 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, 47484756, DOI 10.1016/j.polymer.2014.07.041. (29) Burgess, S. K.; Kriegel, R. M.; Koros, W. J. Carbon dioxide sorption and transport in amorphous poly(ethylene furanoate). Macromolecules 2015, 48, 2184-2193, DOI 10.1021/acs.macromol.5b00333. (30) Eerhart, A. J. J. E.; Faaij, A. P. C.; Patel, M. K. Replacing fossil based PET with biobased PEF: Process analysis, energy and GHG balance. Energy & Environ. Sci. 2012, 5, 6407-6422, DOI 10.1039/c2ee02480b. (31) Burgess, S. K.; Mikkilineni, D. S.; Yu, D. B.; Kim, D. J.; Mubarak, C. R.; Kriegel, R. M.; Koros, W. J. Water sorption in poly(ethylene furanoate) compared to poly(ethyleneterephthalate). Part 2: Kinetic sorption. Polymer 2014, 55, 6870-6882, DOI 10.1016/j.polymer.2014.10.065. (32) Papageorgiou, G. Z.; Papageorgiou, D. G.; Tsanaktsis, V.; Bikiaris, D. N. Synthesis of the bio-based polyester poly(propylene 2,5-furan dicarboxylate), comparison of thermal behavior and solid state structure with its terephthalate and naphthalate

homologues.

Polymer

2015,

62,

28-38,

DOI

10.1016/j.polymer.2015.01.080. (33) Wu, L.; Mincheva, R.; Xu, Y.; Raquez, M. J.; Dubois, P. High molecular weight poly(butylene succinate-co-butylene furandicarboxylate) copolyesters: from catalyzed polycondensation reaction to thermomechanical properties. Biomacromolecules, 2012, 13, 2973-2981, DOI 10.1021/bm301044f. 29

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(34) Zhou, W. D.; Wang, X. W.; Yang, B.; Xu, Y.; Zhang, W.; Zhang, Y. J.; Ji, J. H. Synthesis, physical properties and enzymatic degradation of bio-based poly(butylene adipate-co-butylene furandicarboxylate) copolyesters. Polym. Degrad. Stab. 2013, 98, 2177-2183, DOI 10.1016/j.polymdegradstab.2013.08.025. (35) Andreia, F. S.; Marina, M.; Carmen, S. R. F.; Armando, J. D. S.; Jorge, F. J. C. New copolyesters derived from terephthalic and 2, 5-furandicarboxylic acids: a step forward in the development of biobased polyesters. Polymer 2013, 54, 513-519, DOI 10.1016/j.polymer.2012.11.081. (36) Hu, H.; Zhang, R. Y.; Shi, L.; Ying, W. B.; Wang, J. G.; Zhu, J. Modification of poly(butylene 2,5-furandicarboxylate) with lactic acid for biodegradable copolyesters with good mechanical and barrier properties. Ind. Eng. Chem. Res. 2018, 57, 1102011030, DOI 10.1021/acs.iecr.8b02169. (37) Vijjamarri, S.; Streed, S.; Serum, E. M.; Sibi, M. P.; Du, G. D. Polymers from bioderived resources: synthesis of poly(silylether)s from furan derivatives catalyzed by a Salen Mn(V) complex. ACS Sustainable Chem. Eng. 2018, 6, 2491-2497, DOI 10.1021/acssuschemeng.7b03932. (38) Kim, T.; Koo, J. M.; Ryu, M. H.; Jeon, H.; Kim, S. M.; Park, S. A.; Oh, D. X.; Park, J.; Hwang, S. Y. Sustainable terpolyester of high Tg based on bio heterocyclic monomer of dimethyl furan-2,5-dicarboxylate and isosorbide. Polymer 2017, 132, 122132, DOI 10.1016/j.polymer.2017.10.052. (39) Chi, D.Q.; Liu, F.; Na, H. N.; Chen, J.; Hao, C. 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 Sustainable Chem. Eng. 2018, 6, 9893-9902, DOI 10.1021/acssuschemeng.8b01105. (40) Wang, J. G.; Liu, X. Q.; Jia, Z.; Sun, L. Y.; Zhang, Y. J.; Zhu, J. Modification of poly(ethylene 2,5-furandicarboxylate) (PEF) with 1,4-cyclohexanedimethylene diol: influence of stereochemistry of 1,4-cyclohexylene units. Polymer 2018, 137, 173-185, DOI 10.1016/j.polymer.2018.01.021. (41) Kim, J. H.; Kim, J. R.; Ahn, C. H. Novel biobased copolyesters based on 1,2propanediol or 2,3-butanediol with the same ethylene skeletal structure as PETG. 30

ACS Paragon Plus Environment

Page 30 of 34

Page 31 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Polymer 2018, 135, 314-326, DOI 10.1016/j.polymer.2017.12.024. (42) Terzopoulou, Z.; Kasmi, N.; Tsanaktsis, V.; Doulakas, N.; Bikiaris, D. N.; chilias, D. S.; Papageorgiou, G. Z. Synthesis and characterization of bio-based polyesters: poly(2-methyl-1,3-propylene-2,5-furanoate), poly(isosorbide-2,5furanoate), poly(1,4-cyclohexanedimethylene-2,5-furanoate). Materials 2017, 10, 801, DOI 10.3390/ma10070801. (43) Pinnau, I.; Morisato, A.; He, Z. Influence of side-chain length on the gas permeation properties of poly(2-alkylacetylenes). Macromolecules, 2004, 37, 28232828, DOI 10.1021/ma0498363. (44) 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, 13831391, DOI 10.1021/ma5000199. (45) Soccio, M.; Costa, M.; Lotti, N.; Gazzano, M.; Siracusa, V.; Salatelli, E.; Manaresi, P. Novel fully biobased poly(butylene 2,5-furanoate/diglycolate) copolymers containing ether linkages: Structure-property relationships. Eur. Polym. J. 2016, 81, 397-412, DOI 10.1016/j.eurpolymj.2016.06.022. (46) Guidotti, G.; Soccio, M.; Siracusa, V.; Gazzano, M.; Salatelli, E.; Munari, A.; Lotti, N. Novel random PBS-based copolymers containing aliphatic side chains for sustainable

flexible

food

packaging.

Polymers

2017,

9,

724,

DOI

10.3390/polym9120724. (47) Guidotti, G.; Soccio, M.; Siracusa, V.; Gazzano, M.; Salatelli, E.; Munari, A.; Lotti, N. Novel random copolymers of poly(butylene 1,4-cyclohexane dicarboxylate) with outstanding barrier properties for green and sustainable packaging: content and length of aliphatic side chains as efficient tools to tailor the material’s final performance. Polymers 2018, 10, 866, DOI 10.3390/polym10080866. (48) Montaudo, G.; Puglisi, C.; Samperi, F. Primary thermal degradation mechanisms of PET and PBT. Polym. Degrad. Stab. 1993, 42, 13-28, DOI 10.1016/01413910(93)90021-A. (49) Fang, C.; Chen, T. Advances in civil environmental and materials research (ACEM’ 31

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 34

12). 2012, Korea, August 26-30. (50)Tsanaktsis, V.; Vouvoudi ,

E.; Papageorgiou, G. Z.; Papageorgiou, D. G.;

Chrissafis, K.; Bikiaris, D. N. Thermal degradation kinetics and decomposition mechanism of polyesters based on 2,5-furandicarboxylic acid and low molecular weight aliphatic diols. J. Anal. Appl. Pyrol. 2015, 112, 369-378, DOI 10.1016/j.jaap.2014.12.016. (51)Terzopoulou, Z.; Tsanaktsis, V.;

Nerantzak, M.; Achilias, D. S.; Vaimakis, T.;

Papageorgiou, G. Z.; Bikiaris, D. N. Thermal degradation of biobased polyesters: kinetics and decomposition mechanism of polyesters from 2,5-furandicarboxylic acid and long-chain aliphatic diols. J. Anal. Appl. Pyrol. 2016, 117, 162-175, DOI 10.1016/j.jaap.2015.11.016. (52) Soccio, M.; Lotti, N.; Finelli, L.; Gazzano, M.; Munari, A. Neopenthyl glycol containing poly(propylene terephthalate)s: structure-properties relationships. J. Polym.r Sci.: Part B: Polym. Phys. 2008, 46,170-181, DOI 10.1002/polb.21352. (53) Soccio, M.; Lotti, N.; Finelli, L.; Gazzano, M.; Munari, A. Neopenthyl glycol containing poly(propylene azelate)s: synthesis and thermal properties. Eur. Polym. J. 2007, 43, 3301-3313, DOI 10.1016/j.eurpolymj.2007.06.011. (54) Lee, J. S.; Leisen, J.; Choudhury, R. P.; Kriegel, R. M.; Beckham, H. W.; Koros, W. J. Antiplasticization-based enhancement of poly(ethylene terephthalate) barrier properties. Polymer 2012, 53, 213-222, DOI 10.1016/j.polymer.2011.11.006. (55) Wijmans, J. G.; Baker, R. W. The solution-diffusion model: a review. J Membr. Sci. 1995, 107, 1-21, DOI 10.1016/0376-7388(95)00102-i. (56) Zekriardehani, S.; Jabarin, S. A.; Gidley, D. R.; Coleman, M. R. Effect of chain dynamics, crystallinity, and free volume on the barrier properties of poly(ethylene terephthalate) biaxially oriented films. Macromolecules 2017, 50, 2845-2855, DOI 10.1021/acs.macromol.7b00198. (57) Light, R. R.; Seymour, R. W. Effect of sub-Tg relaxations on the gas transport properties

of

polyesters.

Polym.

Eng.

Sci.

1982,

22,

857-864,

DOI

10.1002/pen.760221402. (58) Genovese, L.; Soccio, M.; Lotti, N.; Munari, A.; Szymczyk, A.; Paszkiewicz, S.; 32

ACS Paragon Plus Environment

Page 33 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Linares, A.; Nogales, A.; Ezquerra, T. A. Effect of chemical structure on the subglass relaxation dynamics of biobased polyesters as revealed by dielectric spectroscopy: 2,5-furandicarboxylic acid vs. trans-1,4-cyclohexanedicarboxylic acid. Phys. Chem. Chem. Phys. 2018, 20, 15696-15706, DOI 10.1039/c8cp01810c. (59) Dennis, J. M.; Enokida, J. S.; Long, T. E. Synthesis and characterization of decahydronaphthalene containing polyesters. Macromolecules 2015, 48, 8733-8737, DOI 10.1021/acs.macromol.5b02288. (60) Kyu, K. B.; Mo, J. H.; Ho, L. Y. Melt rheology of poly(ethylene terephthalate), polyarylate, and their blends. J. Appl Polym. Sci. 1990, 40, 1805-1818, DOI 10.1002/app.1990.070401101. (61) Hu, M.; Xia, Y.; McKenna, G. B.; Kornfield, J. A.; Grubbs, R. H. Linear rheological response of a series of densely branched brush polymers. Macromolecules 2011, 44, 6935-6943, DOI 10.1021/ma2009673. (62) Finkelshtein, E. S.; Makovetskii, K. L.; Gringolts, M. L.; Rogan, Y. V.; Golenko, T. G.; Starannikova, L. E.; Yampolskii, Y. P.; Shantarovich, V. P.; Suzuki, T. Additiontype polynorbornenes with Si(CH3)3 side groups:  synthesis, gas permeability, and free volume. Macromolecules 2006, 39, 7022-7029, DOI 10.1021/ma061215h. (63) Kipnusu, W. K.; Elsayed, M.; Kossack, W.; Pawlus, S.; Adrjanowicz, K.; Tress, M.; Mapesa, E. U.; Krause-Rehberg, R.; Kaminski, K.; Kremer, F. Confinement for more space: A larger free volume and enhanced glassy dynamics of 2-ethyl-1-hexanol in

nanopores.

J.

Phys.

Chem.

Lett.

2015,

6,

3708-3712,

DOI

10.1021/acs.jpclett.5b01533. (64) Zekriardehani, S.; Jabarin, S. A.; Gidley, D. R.; Coleman, M. R. Effect of chain dynamics, crystallinity, and free volume on the barrier properties of Poly(ethylene terephthalate) biaxially oriented Films. Macromolecules 2017, 50, 2845-2855, DOI 10.1021/acs.macromol.7b00198. (65) Mondschein, R.J.; Dennis, J. M.; Liu, H.; Ramakrishnan, R. K.; Nazarenko, S.; Turner, S. R.; Long, T. E. Synthesis and characterization of amorphous bibenzoate copolyesters: permeability and rheological performance. Macromolecules 2017, 50, 7603-7610, DOI 10.1021/acs.macromol.7b01595. 33

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC GRAPHIC

Lateral groups on 1,3-propanediol significantly affect the barrier and crystallization properties of poly(propylene furandicarboxylate)

34

ACS Paragon Plus Environment

Page 34 of 34