Synthesis and Structure–Property Relationship of Biobased

7 days ago - Key Laboratory of Bio-based Polymeric Materials Technology and Application of Zhejiang Province, Ningbo Institute of Materials Technology...
0 downloads 3 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Synthesis and structure–property relationship of bio-based biodegradable poly(butylene carbonate-co-furandicarboxylate) Han Hu, Ruoyu Zhang, Jinggang Wang, Wu Bin Ying, and Jin Zhu ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b00174 • Publication Date (Web): 30 Apr 2018 Downloaded from http://pubs.acs.org on May 5, 2018

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 30 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

Synthesis and structure–property relationship of bio-based biodegradable poly (butylene carbonate-co-furandicarboxylate) Han Hu, †,‡ Ruoyu Zhang, †* Jinggang Wang, † Wu Bin Ying, † Jin Zhu†* †

Key Laboratory of Bio-based Polymeric Materials Technology and Application of

Zhejiang Province, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Zhongguan West Road 1219, Ningbo 315201, People’s Republic of China ‡

University of Chinese Academy of Sciences, Beijing 100049, People’s Republic of

China Corresponding authors: [email protected] (Ruoyu Zhang); [email protected] (Jin Zhu)

Abstract: A series of bio-based poly(butylene carbonate-co-furandicarboxylate) (PBCF), are synthesized through a two-step polycondensation reaction. Chemical structures, thermal properties, crystallization behaviors, mechanical properties, barrier properties and enzymatic degradation of PBCFs are investigated. The linear variation of the glass transition temperatures with the content ratio confirms the good miscibility between butylene carbonate and butylene furandicarboxylate units. BF segments could crystallize in most contents under different temperatures. Consequently, mechanical properties of these copolymers not only depend on the composition but also on the annealing conditions. Long time annealing at room temperature or short time annealing under high temperature could tremendously

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

increase the tensile modulus. For room temperature annealing, the formation of less perfect crystals of the PBF could interpret the enhancement of modulus. On the other hand, the high temperature annealing induces more perfect PBF crystals and improves the modulus significantly too. The tensile toughness of the PBCFs is good with the lowest elongation at break of 260%, and the degradation could be observed in samples with less than 60 mol% BF units. With good gas barrier properties, fast biodegradability and high mechanical performance, these copolyesters own potential applications in plastic industry. Keywords: Bio-based polymer, 2,5-Furandicarboxylic acid, Poly (ester carbonate),

Mechanical Property, Biodegradability, Gas barrier.

Introduction 2,5-Furandicarboxylic acid (FDCA) is one of the 12 potential platform chemicals selected by the U.S. Department of Energy derived from biomass. The world’s leading chemical companies, such as Avantium, DuPont, BASF, Archer Daniels Midland and Coca-Cola, show strong interest on specific FDCA based polymers and solidly promote the related technologies and usability as engineering, packaging, fiber forming and elastomeric materials.1-2 FDCA has been regarded as one of the most high-potential bio-based aromatic monomers.3-5 It is a perfect bio-based alternative to the petroleum-based terephthalate acid (TPA), even if less aromatic but more rigid.6-7 FDCA based aromatic polyesters poly(ethylene furandicarboxylate) (PEF),8-13 poly(propylene furandicarboxylate) (PPF)9, 14-16 and poly(butylene furandicarboxylate) (PBF)7,

17-19

have gained great interest because of their excellent mechanical and

ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30 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

barrier properties. However, the aromatic polyesters are non-degradable, recycling will be more complicated when plastics have been mixed with organic fractions.20-21 As the pollution caused by plastic waste becomes more and more serious, biodegradable polymers become more important to ensure the sustainability of the plastic industry and human society. Therefore, many papers have been published concerning the biodegradability of FDCA based copolyesters, such as poly(butylene succinate-co-furandicarboxylate),22-23 furandicarboxylate)23-25

and

poly(butylene

poly(butylene

adipate-co-

sebacate-co-furandicarboxylate).26

Aliphatic diacids (succinic acid, adipic acid, sebacic acid) were used but the elastic modulus and tensile strength of copolyesters were relatively low and the barrier properties were poor on account of the introduction of aliphatic diacids. Some aliphatic polyesters,27-35 such as poly(lactic acid), poly(butylene succinate) and polycaprolactone show good biodegradability and attract lots of attention in the past several decades. Recently, aliphatic polycarbonates (APCs),36-40 as new biodegradable polymers, draw lots of attention due to their nontoxicity, biodegradability, and biocompatibility. One of the best strategy for preparing aliphatic polycarbonates is the condensation polymerization of dimethyl carbonate (DMC) and aliphatic diols. DMC is has been realized mass production using either carbon monoxide or carbon dioxide, and the cost is lower than the ring-opening polymerization. It is nontoxic and the carbonate linkages are susceptible to hydrolytic cleavage, which makes it a truly green material.41

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

If we can introduce proper amount of DMC units into FDCA based polymer, then the resulted copolyesters may possess the advantages of both species. Unlike long chain aliphatic diacids, the introduction of DMC will not only promote the biodegradability of copolymers, but also keep the rigidity of the chain segment, showing relatively high barrier and mechanical performance. In this paper, a series of PBCF copolymers with BF content from 30% to 90% are synthesized. Basic properties of these polymers are investigated, including the chemical structures, thermal properties, crystal structures, thermal stability and mechanical properties, barrier performances and degradation ability. Moreover, the structure-property relationship was studied by correlating the thermal history or crystallinity with the mechanical properties. Isothermal annealing experiments and room temperature storage can promote the crystallinity and further alter the mechanical properties of PBCFs. With well performances in many aspects, PBCFs could be developed further for the next generation of materials with good mechanical and barrier properties and excellent biodegradability.

Experiments Materials. 1,4-Butanediol(BDO, 99%),dimethyl carbonate(DMC, 98%),tetrabutyl titanate (TBT, ≥99%)and antimony trioxide(Sb2O3, 99%)were purchased from Aladdin Reagent Co. Ltd (Shanghai, China). 2,5-furandicarboxylic acid (FDCA, 99.9%) was purchased from Chem target Technologies Co., Ltd. (Mianyang China). PBAT was purchased from Jinhui Technologies Co., Ltd. (Shanxi China). PLA (4032D) was purchased from Natureworks.

ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 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

Synthesis of Dimethyl Furan-2,5-dicarboxylate (DMFD). The synthesis of DMFD followed the procedures described in our previous literature.13 Its chemical structure was confirmed by 1H NMR and 13C NMR. 1

H NMR (400 MHz, DMSO-d6, ppm): 3.88 (6H), 7.41 (2H).

13

C NMR (400 MHz, DMSO-d6, ppm): 52.9 (2C), 119.6, 146.7 (4C), 158.7 (2C).

Synthesis of PBCFs. As shown in Scheme 1, the copolymers were synthesized by an unconventional method. Since the DMC has a low melting temperature of 90 °C, it could volatile rigorously at the transesterification temperature of DMFD and BDO (180 °C). Therefore, two kinds of oligomers were prepared firstly.

Scheme 1. Synthesis of poly(butylene carbonate-co-furandicarboxylate) For the synthesis of oligo(butylene carbonate) (OBC), DMC and BDO at a molar ratio of 2.5:1 with catalyst of TBT, were added into the 1L three-necked round bottom flask

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

and then were heated to 85-90 °C under nitrogen atmosphere. 1H-NMR was used to monitor the reaction progress. When more than 95% hydroxyl group of BDO reacted with DMC, the reaction temperature was gradually increased to 180 °C in 3 h and maintained for 1 h to ensure that the unreacted DMC and methanol could be removed completely. The oligo(butylene furandicarboxylate) (OBF) was synthesized by a conventional transesterification step. BDO, DMFD (BDO/DMFD=1.6/1) and TBT were added into a 1L three-necked round bottom flask equipped with a mechanical stirrer with torque indicator, a N2 inlet and a condenser. The flask was evacuated to less than 50 Pa and then purged with high purity N2 (99.999%) for three times to ensure a total N2 atmosphere. The reaction was conducted at 170 °C for 1 h and 180 °C for about 3-5 hours. The transesterification was considered to be completed if more than 95% of the theoretical methanol was distilled out. Poly(butylene carbonate) (PBC) and poly(butylene furandicarboxylate) were homopolyesters and were prepared by the melt polycondensation of OBC and OBF, respectively. For PBCFs, different OBC and OBF ratios were chosen to get the targeted copolymers with BF content ranged from 30% to 90%. Appropriate amounts of oligomers and Sb2O3 were added into the reactor flask. The reaction temperature was controlled at 210 °C to avoid the thermal degradation of DMC group. The pressure of the system was gradually decreased to 10-30 Pa in about 0.5 h to minimize oligomer sublimation. As the reaction went on, the rotating speed was reduced from 180 rpm to 30 rpm. The reaction stopped when the maximal torque value of the stirrer

ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 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

kept unchanged. Then the system pressure was restored to normal atmospheric pressure by the inlet of N2 and the target products were obtained.

Characterization All of the characterization methods were summarized in supporting information.

Result and discussion The chemical structures and compositions of PBC, PBF and PBCFs.

Figure 1. 1H NMR spectra of PBC, PBF and PBCF. (a) The whole spectra; (b) magnification of chemical shifts a; (c) magnification of chemical shifts b; (d) corresponding chemical structures of FBF, FBC, CBC units in PBCFs. The chemical structures of the samples were analyzed with ATR-FTIR (Figure S1) and 1H NMR (Figure 1). The signal of CH in the furan rings located at 7.23 ppm (c). CH2 near the ester bonds in BDO unit was at 4.17 (a4), 4.20 (a3), 4.37 (a2) and 4.40

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 8 of 30

(a1) ppm (Figure 1b), while the middle CH2 in BDO unit was at 1.77 (b4), 1.82 (b2), 1.86 (b2) and 1.92 (b1) ppm (Figure 1c). PBC and PBCF30 to 70 could be dissolved in CDCl3, while PBCF80, PBCF90 and PBF had to be dissolved in CF3COOD. As a result, the chemical shifts became slightly different in the two kinds of solvents (Figure 1b and 1c). The content of furanic unit (ϕBF) could be calculated by using eqs. 1, where Ia1, Ia2, Ia3, Ia4 were the integral intensities of the CH2 in BDO unit that were close to ester bonds, and Ic was the integral intensities of the CH in furan ring. It could be seen that the peak areas of a1 and b1 would decrease and those of a4 and b4 could increase with decreasing ϕBF.

 =



, = 1 + , = 1 + R=



(1)



(2)

   

 ,

+

   

(3)

  

(4)

,

The number-average sequence length of BF and BC units (Ln,BF and Ln,BC), and the degree of randomness (R) were calculated using eqs. 2-4. As shown in Table 1, the PBCFs were random copolymers as R was close to 1. While Ln,BF increased with ϕBF, Ln,BC decreased accordingly. The copolyesters were considered to be biodegradable if the number-average sequence length of aromatic units was less than 3.

42

By such

assumption, since the Ln,BF of PBCF60 was 2.66, PBCF30 to PBCF60 should be biodegradable. In fact, such deduction could be verified by biodegradation experiment, which would be shown later. Table 1. Molecular characteristics and thermal stability of PBC, PBF and PBCF. Sample

BF (mol%)a

Ln,BFb

Ln,BCb

R

[η]c

Mwd

DI

T5%,°C

Td, max, °C

PBC

0

/

/

/

1.15

86,300

2.10

318

373

PBCF30

32

1.54

3.10

0.97

0.95

68,200

1.90

338

397

ACS Paragon Plus Environment

Page 9 of 30 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

PBCF40

40

1.76

2.56

0.96

0.93

64,700

2.03

339

398

PBCF50

53

2.35

1.82

0.98

0.94

67,500

1.98

344

401

PBCF60

61

2.66

1.65

0.99

0.98

71,300

1.89

351

404

PBCF70

67

3.33

1.43

1.02

0.95

68,600

1.86

352

402

PBCF80

80

5.14

1.24

1.01

1.02

75,600

1.91

359

405

PBCF90

91

10.89

1.10

1.00

0.98

71,000

1.81

364

402

PBF

100

/

/

/

1.00

51,800

2.04

372

406

a

Mole fraction of BF units in the polyesters.

b

Number-average sequence length of BF and BC units.

c

Intrinsic viscosity obtained with an Ubbelohde viscometer.

d

Weight-average molecular weight determined by GPC.

The Mw, DI and intrinsic viscosities were summarized in Table 1. The Mw of the obtained polymers ranged from 64,700 to 86,300 g/mol, with dispersity ranged from 1.8 to 2.1. The intrinsic viscosity of the samples were in the range of 0.93-1.15 dL/g, which revealed that relatively high-molecular weight copolyesters with close Mw were successfully synthesized. Thermal properties Thermal analysis was carried out by DSC for PBC, PBF and PBCFs, as shown in Figure 2. The thermal transition temperatures were summarized in Table S1. All the samples for DSC tests were obtained from the reactor flask after the completion of the reaction and stored at the room temperature for one week. PBC was a typical semi-crystalline polymer. It exhibited high melting enthalpy (∆Hm 45.9 J/g) in the first heating scan, and its melting point was 56.8 °C. But no crystallization peak in the cooling process and only small melting peak in the second heating scan could be found, which meant that the crystallization ability of PBC was poor. For PBCF30, there was no melting peak in the first heating scan, and this was because the insertion of furan units completely depressed the crystallization of PBC, and similar

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

phenomenon could be found in other reports of FDCA based copolyesters.43-44 Both PBCF40 and PBCF50 indicated the melting peaks close to 52 °C, which belonged to less perfect crystals of PBF and will be analyzed in detail below.45-46 PBCF50 indicated higher melting peak at 90.5 °C (∆Hm 5.8 J/g), representing more perfect crystal structure of PBF. PBCF60 showed no melting peaks that in this thermal history PBCF60 crystallized slowly. PBCF70 had a low melting peak at 50.3 °C (∆Hm 3.4 J/g) and a higher melting peak at 123.2 °C (∆Hm 5.6 J/g). PBCF80, PBCF90 and PBF had cold crystallization peaks. The Tm and ∆Hm increased from 143.5 to 168.3 °C and 16 to 34.8 J/g, correspondingly. In the cooling scans, PBCF90 had a crystallization peak at 92 °C (∆Hc 3.0 J/g). PBF crystallized faster with a crystallization peak at 102 °C (∆Hc 18.4 J/g). Other samples showed no crystallization peak in the cooling process. The second heating scans confirmed poor crystallization ability of PBCFs. From PBCF30 to 70, neither cold-crystallized peak nor melting peak appeared. However, they exhibited weak melting peak in the first heating scan, suggesting that they could crystallize under certain conditions. The cold crystallization and melting peaks only appeared at high ϕBF samples. PBCF80 exhibited weak cold crystallization temperature (∆Hcc 4.1 J/g) at 112.7 °C and melting peak at 145.2 °C (∆Hm 4.1 J/g). PBCF90 exhibited relatively stronger cold crystallization at 102.4 °C (∆Hcc 17.7 J/g) and melting peak at 159.3 °C (∆Hm 20 J/g). As can be expected, PBF had the strongest crystallization ability. It was a semicrystalline polyester with the best crystallizability among the series, and exhibited cold crystallization (∆Hcc 23.6 J/g) at 105.0 °C and melting peak (∆Hm 34.2 J/g) at 168.8 °C. Obviously, the crystallization ability of PBCFs decreased with increasing ϕBC.

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30 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 2. DSC curves of (a) the first heating scans, (b) the second heating scans and (c) second heating scans of PBC, PBF and PBCF. In addition, all of PBCFs and PBF had only one Tg as measured by DSC. The Tg increased monotonously with the ϕBF from -32.0 °C (PBC) to 36.1 °C (PBF). The composition dependent Tg fit well with the Fox equation (eqs. 5) as proved in Figure S2. The BC weight fraction and BF weight fraction were abbreviated as ϕω,BC and ϕω,BF.

1 , , = + 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

It seems that the short BC and BF segments have very good thermodynamic compatibility in random copolymers, which maybe induced by the limited segmental length. The crystallization half-time (t1/2) was measured to evaluate the crystallization ability by isothermal experiments. Note that Tg and Tm of these samples were very different, we chose three crystallization temperatures for each sample to compare the crystallization rate at certain temperature range. The crystallization rate drastically increased with increasing ϕBF, which could be ascribed to longer BF sequence. PBF showed the t1/2 of 3.8 min at 110 °C, while PBCF80 and PBCF90 decreased to 7 min and 15 min, respectively. For PBCF60 and PBCF70, their half crystallization time increased to 98 min and 52 min at 80 °C, respectively. As to PBCF50, its t1/2 at 50 °C was as long as 160 min. For PBCF40, its t1/2 was even longer than 200 min and was not shown in the Figure 3. Furthermore, thermomechanical properties of these samples were tested by DMA (Figure S3) and discussed in supporting information.

ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30 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. Crystallization half-time (t1/2) for PBCFs and PBF measured from isothermal experiments at different temperatures. Thermal stability The thermal stability of polymers was a crucial factor for their synthesis, processing, and application. TGA and DTG curves of PBC, PBF and PBCFs were presented in Figure S4. The characteristic decomposition temperatures at 5% weight loss, T5%, and the decomposition temperature at the maximum rate, Td,max, were listed in Table 1. The PBC showed relatively poor thermal stability with T5% less than 320 °C. All the PBCFs had better thermal stability with T5% higher than 330 °C. With increasing content of ϕBF, the T5% increased from 338 °C (PBCF30) to 372 °C (PBF). Such results suggested that the introduction of BF units into PBC chain hindered the thermal degradation of PBC to some extent. The Td,max, on the other hand, were in a narrow range from 397 °C (PBCF30) to 406 °C (PBF), but were far higher than the Td,max of PBC (373 °C). In conclusion, the PBCFs had good thermal stability and could satisfy the daily application. Thermal history dependent mechanical properties. Because of the relatively low glass transition temperatures of copolymers, these samples could undergo slow crystallization when placed at room temperature. It was also widely known that crystallization could affect the mechanical properties a lot. Therefore, it was important to clarify the thermal history of the samples when evaluating their mechanical properties. In this study, three different thermal treatments were applied in each sample to investigate the annealing effect on mechanical

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

properties. The first way was to store the sample at the room temperature for one week after the press molding. The second method was to put the copolymers under room temperature for 4 weeks. Without air condition, the room temperature was in the range of 25-32 °C. From PBCF40 to 70, their Tg were lower than the room temperature. On the other hand, the Tg of PBCF80 and PBCF90 were close to room temperature. As to PBF, it showed Tg higher than the room temperature. The last one was to anneal the samples for 3 hours at elevated temperatures. PBCF80, PBCF90 and PBF were annealed at 110 °C. PBCF60 and PBCF70 were kept at 80 °C, and PBCF50 and PBCF40 were placed under 45 °C. Then, they were cooled down to room temperature naturally. The results of t1/2 was the reference when we chose the annealing temperatures.

ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 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. XRD patterns and DSC curves of PBCFs and PBF with three different thermal treatments; (a), (d) room temperature annealing for one week; (b), (e) room temperature annealing for four weeks; (c), (e) high temperature annealing four 3h. The XRD patterns confirmed the change from nearly amorphous to obviously semicrystalline state after the storage at room temperature or annealing at elevated temperatures. After one week annealing at room temperature, no obvious diffraction

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

peak could be identified in PBCFs and PBF. Nevertheless, in Figure 4d, the heating scan with ramp rate of 20 °C /min in DSC showed melting peaks in PBCF40 and PBCF50. When the room temperature annealing increased to four weeks, clear diffraction peaks could be identified in Figure 4b at 2θ values of 18.1 °, 22.5 ° and 25.0 ° for all the samples except PBF, and these lattice plane belonged to the crystals of PBF. The main reflections of PBC crystal lattice, appeared at 2θ values of 21.2 ° and 21.6 ° (Figure 4a), could not be identified here and implied that there was no PBC crystal in the copolymers. The melting peaks became more obvious in DSC heating scan in Figure 4e. The melting peaks around 60 °C in Figure 4e belonged to less perfect PBF crystals, as has been reported elsewhere.45-46 On the other hand, melting peaks higher than 70 °C belonged to the more perfect PBF crystals. Like shown in Table S2, by subtracting the cold crystallization enthalpy, we could obtain the crystallinity of PBF after annealing. Apparently, the crystallinity of low melting temperature PBF crystal (XC,L) and high melting temperature PBF crystal (XC,H) increased significantly when compared with those after one week’s storage. This indicated that in a longer period of time, crystallization did happen in PBCF40 to 90 at room temperature. However, there was no obvious change of PBF at room temperature, because its Tg was higher than the room temperature. Another important fact was that the Tm of PBF crystals systematically increased with the increase of BF content. Interestingly, annealing at high temperature was somewhat different from the process happened at ambient temperature. XRD profiles in Figure 4c displayed much sharper peaks of PBF crystals for PBCFs. The melting enthalpy of PBF crystals after

ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 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

annealing at elevated temperature was much higher than those annealed at room temperature. With the above observation, it can be said that PBF lamellar crystal with more perfect structure formed during annealing at elevated temperatures in high ϕBF samples. At the same time, when ϕBF was higher than 50%, double melting peaks of PBF crystals appeared in Figure 4f. At the same time, the low melting peaks around 50-60 °C disappeared in every sample. But new melting peaks at 70-120 °C appeared for PBCF50 to 90 and PBF. It was of high possibility that the thickness of PBF lamellar crystal became thicker during the high temperature annealing and their corresponding melting points moved to higher temperatures.

47

Until PBCF90, with

the increase of ϕBF, the difference between the two melting temperatures became increasingly obvious. However, the difference in the two peaks in PBF became less distinct than that of PBCF90. Such phenomenon may be related with the confined diffusion of BF segments in PBCFs and the secondary nucleation and growth phenomenon in all samples. Nevertheless, the mechanism of the double melting peaks of PBF crystals was unclear and required more work in the future. PBC crystal didn’t exist in any of these procedures. One providence was that there was no XRD diffraction peak of PBC crystal. Another was that the melting temperature of PBC was 56 °C and the PBC segments in copolyesters should show even lower Tm. But we didn’t observe such phenomenon. In summary, after high temperature annealing, the low temperature melting peaks at 50-60 °C shifted to higher temperature, the values of XC,L decreased to 0 for all the samples, while XC,H witnessed a gradual increase from 4.7% (PBCF40) to 33.6% (PBF).

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

Mechanical properties of copolymers remarkably varied with the annealing time and conditions. Typical stress-strain curves were shown in Figure S5 for all the samples with different treatments, and the elastic modulus (E), tensile strength (σb) and strain at break (εb) of PBCFs and PBF were compared in Figure 5 and Table S3. Due to the low Tg and poor crystallization ability, the specimen of PBCF30 for mechanical test was difficult to prepare and its mechanical property was not reported in this paper. For the specimens with one week’s room temperature annealing, the elastic modulus of PBCF60 and PBCF70 were less than those of PBCF40 and PBCF50. Such results could be explained by the low PBF crystallinity in PBCF60 and PBCF70, as shown in Table S2. With the continuous increase of rigid BF segment, Tg was higher than testing temperature for PBCF80, PBCF90 and PBF. Therefore, the modulus increased rapidly. After four weeks’ annealing, the elastic modulus of all PBCFs increased. Especially for PBCF60 and PBCF70, their modulus increased more than 10 times from 18.0 to 204.1 MPa and 18.6 to 316.1 MPa, respectively. However, since the Tg of PBF was higher than the room temperature, its modulus remained the same as before. Compared with the room temperature annealing, high temperature treatment showed different effects on the modulus. From PBCF40 to 70, their elastic modulus were even lower than those after four weeks’ annealing. Such phenomenon may be correlated with the crystal density in the samples. Higher crystallization temperature would depress the nucleation density. In contrary, for high ϕBF containing samples, PBCF80, PBCF90 and PBF, their modulus after high temperature annealing showed remarkable increase. Such increment could be attributed to the perfection and increase

ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30 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

of high melting temperature PBF crystals. It seemed that for samples from PBCF40 to 70, low melting temperature PBF crystal was more important than high melting temperature PBF crystal in determining the modulus, while for samples from PBCF 80 to PBF the latter one was dominant. Except for PBCF40, the tensile strength gradually increased with the content of BF in most samples. As to the elongation at break, with annealing treatment it systematically decreased with increasing ϕBF. Nevertheless, even the lowest values of PBCF90 could reach 260%.

Figure 5. Elastic modulus, strength at break, elongation at break of PBCFs and PBF with conditions of room temperature annealing for one week (a), four weeks (b) and high temperature annealing for 3 h (c).

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

Stretch induced crystallization in PBCFs and PBF In our previous result, we found that stretch induced crystallization could occur in PBF.48 In this work, stretch (break at the stretching speed of 20 mm/min) induced crystallization occurred in both PBCFs and PBF. We chose samples after four weeks’ annealing under room temperature for stretch experiment. The XRD patterns and DSC scans before and after stretch were summarized in Figure S6. It was obvious that the characteristic diffraction peaks of PBF at 18.1° and 25.0° became much sharper after stretch. But we noticed that the lattice plane was strengthened differently in different samples. For example, for PBCF40 to 60, the intensity of the diffraction peak at 18.1° was enhanced. In contrary, for samples PBCF80-90 and PBF, the diffraction peak at 25.0° was strengthened. The most interesting case was PBCF70. Before stretch, its XRD peak at 25.0° was strong and the one at 18.1° was not that sharp. However, the situation reversed after the stretch experiment. It seemed that from PBCF40 to 70, the lattice plane at 18.1° became more perfect after the stretch. Nevertheless, as to the samples with BF content higher than 70, only the structure at 25.0° could be reinforced. It was a very interesting phenomenon and worth or more experimental work in the future. Besides, we could see that the melting enthalpy of both low and high melting temperature PBF crystals increased after elongation. The XC,L and XC,H, before and after stretch, were calculated and listed in Table S4. For PBF, the stretch induced crystallization was the most prominent. It evolved from a completely amorphous state to a semicrystalline state. As to PBCF40, the stretch didn’t induce the formation of high melting temperature crystal. While for PBCF90, the low melting

ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 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

temperature crystal could grow during the stretch. It was also interesting to see that the cold crystallization of PBF disappeared after the stretch, implying that the crystal developed well during the elongation procedure. The stretch induced crystallization could enhance the strength of the materials during the elongation, and may benefit in toughening application.48 Barrier properties Table 2. CO2, O2 and H2O barrier properties of PBCFs and PBF. Sample a

CO2

BIFp

(barrer) b

O2

H2O

BIFp

(barrer) c

BIFp

(g·cm/cm2·s·Pa) d

PBAT

5.9

1

0.76

1

3.52*10-13

1

PLA

1

5.9

0.25

3.0

1.1*10-13

3.2

PBCF40

0.70

8.4

0.10

7.6

1.30*10-13

2.7

PBCF50

0.30

19.7

0.046

16.5

6.6*10-14

5.3

PBCF60

0.25

23.6

0.037

20.3

4.3*10-14

8.2

PBCF70

0.22

26.8

0.025

30.4

3.9*10-14

9.0

PBCF80

0.19

31.0

0.020

38.0

3.5*10-14

10.0

PBCF90

0.16

36.8

0.015

50.6

3.0 *10-14

11.7

PBF

0.08

73.7

0.012

63.3

2.3*10-14

15.3

a

The test was performed at low pressure (0.1001 MPa).

b

Carbon dioxide permeability coefficient, at 23 °C, 50% relative humidity. 1 barrer=

10-10 cm3 cm/cm2·s·cm Hg. c

Oxygen permeability coefficient, at 23 °C, 50% relative humidity.

d

Water vapor transmission rate, at 38 °C, 90% relative humidity.

Concerning the applications like packaging and agricultural films, the gas barrier properties were of great concern. For most biodegradable materials like PLA, and PBAT, their gas barrier properties were relatively poor, which limited further usage in

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

related fields. In this study, the CO2 and O2 permeability coefficient were investigated by the complementary pressure decay sorption and permeation techniques, and the results were summarized in Table 2. Gas barrier properties of commercial products of PBAT and PLA were also measured. The Barrier Improvement Factor (BIFp) represented the gas permeability coefficient of PBAT divided by the permeability of PLA, PBCFs and PBF. Like reported by other researchers,49-50 the furan structure could significantly improve the gas barrier properties of polyesters due to its asymmetric rigid structure and the polarity of the furan rings.51 As evidenced by PBCF40 and PBAT, which had similar aromatic unit ratio in the main chain, the CO2 BIFp of PBCF40 could be as high as 8.4. When ϕBF increased, the CO2 BIFp kept on increasing and reached 73.7 of PBF. Similarly, as to O2 permeability coefficient, the O2 BIFp of PBCFs increased from 7.6 of PBCF40 to 63.3 of PBF. The O2 barrier property of PBCFs and PBF was also enhanced substantially when compared with PBAT and PLA. As for H2O barrier property, most of PBCFs showed improved barrier ability than PBAT and PLA too. The WTR (water vapour transmission rate) showed similar trend but relatively slower growth degree than gas barrier properties. The H2O BIFp was 2.7 for PBCF40 and 5.3 for PBCF50, and then it increased to 11.7 for PBCF90 and 15.3 for PBF. The H2O BIFp of PLA was 3.2, which located between that of PBCF40 and PBCF50. The low permeability could be attributed to hindered rotation and polar feature of the furan rings.

49

Moreover, using of DMC to replace

adipic acid as a fatty diacid monomer could increase the rigidity of the main chain too.

ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 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

Biodegradation of PBCFs. Since the white pollution became more and more serious, biodegradation was also an attractive property in packaging and film materials. A good combination of biodegradation with mechanical properties was highly desirable. In this paper, porcine pancreas lipase was selected as the enzyme to decompose the PBCFs and PBF. The biodegradation experiment was carried out at 37 °C and in a phosphate buffered saline (PBS) solution with pH=7.4. Samples were weighed at different soak times and the weight loss was recorded. As can be found in Figure 6, samples from PBCF30 to 60 were biodegradable. PBCF30 showed a biodegradation rate with a weight loss of 14.5% after 28 days, while PBCF40 had a relatively slower biodegradation rate of 10.4%. The degradation rates of PBCF50 and PBCF40 were 6.3% and 4.2%, respectively. The introduction of furan units into PBC weakened the biodegradability, and the degradation phenomenon disappeared in samples with BF content higher than 60%.

Figure 6. The weight loss curves of PBCFs and PBF during enzymatic degradation.

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

The compositions of PBCFs before and after enzymatic degradation were tested by 1H NMR. As summarized in Table S5, the degradation behavior led to a slight increase of ϕBF, suggesting that more BC units were broken during degradation. Besides, the weight average molecular weight and intrinsic viscosity after degradation were shown in Table S5. The weight average molecular weight and intrinsic viscosity decreased significantly for PBCF30 to 60, which solidly improved the degradation happened in the whole sample.

Conclusion A series of PBCFs have been successfully synthesized by a two-step melting polymerization. 1H NMR results confirm the randomness of the chemical unit distribution along the main chain, while the linear variation of Tg with composition ratios proved the good miscibility between BC and BF structures. The BC component could not crystallize in the copolymer, and BF units could form two kinds of crystals with low and high melting temperature, respectively. The diffraction peaks of PBF crystal lamellar is clearly shown in XRD curves, especially after the treatment of proper annealing or stretch. The mechanical properties of PBCFs depended largely on the crystallization state of PBF. With long time annealing at room temperature, the increased low melting temperature PBF crystals improved the modulus a lot. On the other hand, after annealing at high temperature, the high melting temperature PBF crystals are well developed and tremendously increase the tensile modulus. Nevertheless, the modulus of PBCFs tend to increase with the content of BF component. At the same time, the tensile strength increase in a more moderate way,

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30 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

and the elongation at break decreases systematically. Just like pure PBF, PBCFs exhibit stretch induced crystallization phenomenon. Moreover, by changing the component in the main chain, we could adjust the degradation speed. In consequence, with good elongation at break and acceptable mechanical properties, these biodegradable copolymers could be further applied in blending with other biodegradable polymers like PLA, PBAT etc. to achieve polymer blends with balanced properties. The good thermal stability under 300 °C could ensure the processability in plastic industry. In conclusion, this paper provides us some FDCA-based copolymers with increased thermal stability, higher melting points, good mechanical properties, better barrier properties and adjustable degradation rate.

Supporting Information Characterization methods; ATR-FTIR spectra; Composition-dependence Tg of PBCFs as a function of BF weight fraction; TGA and DTG traces; DMA spectra; Stress-strain curves, mechanical properties and crystallinity with three different thermal history; XRD patterns, DSC scans and crystallinity before and after stretch. Mole fraction, weight average molecular weight and intrinsic viscosity before and after enzymatic degradation.

Acknowledgements This work is supported by National Key Research and Development Program of China (Grant No. 2017YFB0303000) and National Natural Science Foundation of China (Grant No. 51773218). We thank Prof. Chuncheng Li in Institute of Chemistry, Chinese Academy of Sciences in the synthesis of PBC homopolymer.

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 26 of 30

References 1.

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. Polymer Chemistry 2015, 6 (33), 5961-5983. 2.

Vilela, C.; Sousa, A. F.; Fonseca, A. C.; Serra, A. C.; Coelho, J. F. J.; Freire, C. S. R.; Silvestre, A.

J. D., The quest for sustainable polyesters - insights into the future. Polymer Chemistry 2014, 5 (9), 3119-3141. 3.

Zhang, J.; Li, J.; Tang, Y.; Lin, L.; Long, M., Advances in catalytic production of bio-based

polyester monomer 2,5-furandicarboxylic acid derived from lignocellulosic biomass. Carbohydrate polymers 2015, 130, 420-428. 4.

Moreau, C.; Belgacem, M. N.; Gandini, A., Recent catalytic advances in the chemistry of

substituted furans from carbohydrates and in the ensuing polymers. Topics in Catalysis 2004, 27 (1-4), 11-30. 5.

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. European Polymer Journal 2016, 83, 202-229. 6.

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 (4), 1383-1391. 7.

Zhu, J.; Cai, J.; Xie, W.; Chen, P.-H.; 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 (3), 796-804. 8.

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(ethylene terephthalate). Part 2: Kinetic sorption. Polymer 2014, 55 (26), 6870-6882. 9.

Wang, J.; Liu, X.; Zhu, J.; Jiang, Y., Copolyesters Based on 2,5-Furandicarboxylic Acid (FDCA):

Effect of 2,2,4,4-Tetramethyl-1,3-Cyclobutanediol Units on Their Properties. Polymers 2017, 9 (9), 305. 10. 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. Physical Chemistry Chemical Physics 2014, 16 (17), 7946-7958. 11. Wang, J.; Liu, X.; Jia, Z.; Sun, L.; Zhang, Y.; Zhu, J., Modification of poly(ethylene 2,5-furandicarboxylate) (PEF) with 1, 4-cyclohexanedimethanol: Influence of stereochemistry of 1,4-cyclohexylene units. Polymer 2018, 137, 173-185. 12. Wang, J.; Liu, X.; Jia, Z.; Liu, Y.; Sun, L.; Zhu, J., Synthesis of bio-based poly(ethylene 2,5-furandicarboxylate) copolyesters: Higher glass transition temperature, better transparency, and good barrier properties. Journal of Polymer Science Part A: Polymer Chemistry 2017, 55 (19), 3298-3307. 13. Wang,

J.;

Liu,

X.;

Zhang,

Y.;

Liu,

F.;

Zhu,

J.,

Modification

of

poly(ethylene

2,5-furandicarboxylate) with 1,4-cyclohexanedimethylene: Influence of composition on mechanical and barrier properties. Polymer 2016, 103, 1-8.

ACS Paragon Plus Environment

Page 27 of 30 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

14. 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. 15. Vannini, M.; Marchese, P.; Celli, A.; Lorenzetti, C., Fully biobased poly(propylene 2,5-furandicarboxylate) for packaging applications: excellent barrier properties as a function of crystallinity. Green Chem. 2015, 17 (8), 4162-4166. 16. Hu, H.; Zhang, R.; Wang, J.; Ying, W. B.; Zhu, J., Fully bio-based poly(propylene succinate-co-propylene furandicarboxylate) copolyesters with proper mechanical, degradation and barrier properties for green packaging applications. European Polymer Journal 2018, 102, 101-110. 17. Ma,

J.;

Yu,

X.;

Xu,

J.;

Pang,

Y.,

Synthesis

and

crystallinity

of

poly(butylene

2,5-furandicarboxylate). Polymer 2012, 53 (19), 4145-4151. 18. Sousa, A. F.; Guigo, N.; Pożycka, M.; Delgado, M.; Soares, J. R.; Mendonça, P.; J. Coelho, J. F.; Sbirrazzuoli, N.; Silvestre, A., Tailor design of renewable copolymers based on poly(1,4-butylene 2,5-furandicarboxylate) and poly(ethylene glycol) with refined thermal properties. Polymer Chemistry 2017, 9 (6), 722-731. 19. Soccio, M.; Martinez-Tong, D. E.; Alegria, 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. 20. Nofar, M.; Tabatabaei, A.; Sojoudiasli, H.; Park, C. B.; Carreau, P. J.; Heuzey, M. C.; Kamal, M. R., Mechanical and bead foaming behavior of PLA-PBAT and PLA-PBSA blends with different morphologies. European Polymer Journal 2017, 90, 231-244. 21. Hua, L.; Chen, Q.; Yin, J.; Zhang, C.; Wang, X.; Yin, J.; Feng, X.; Yang, J., Fabrication and Physical Properties of Poly(ε‐Caprolactone)/Modified Graphene Nanocomposite. Macromolecular Materials & Engineering 2016, 1600328. 22. Wu, L.; Mincheva, R.; Xu, Y.; Raquez, J. M.; Dubois, P., High molecular weight poly(butylene succinate-co-butylene furandicarboxylate) copolyesters: from catalyzed polycondensation reaction to thermomechanical properties. Biomacromolecules 2012, 13 (9), 2973-81. 23. Peng, S.; Wu, B.; Wu, L.; Li, B.-G.; Dubois, P., Hydrolytic degradation of biobased poly(butylene succinate-co-furandicarboxylate) and poly(butylene adipate-co-furandicarboxylate) copolyesters under mild conditions. Journal of Applied Polymer Science 2017, 134 (15). 24. Wu, B.; Xu, Y.; Bu, Z.; Wu, L.; Li, B.-G.; Dubois, P., Biobased poly(butylene 2,5-furandicarboxylate) and poly(butylene adipate-co-butylene 2,5-furandicarboxylate)s: From synthesis using highly purified 2,5-furandicarboxylic acid to thermo-mechanical properties. Polymer 2014, 55 (16), 3648-3655. 25. Zhou, W.; Wang, X.; Yang, B.; Xu, Y.; Zhang, W.; Zhang, Y.; Ji, J., Synthesis, physical properties and enzymatic degradation of bio-based poly(butylene adipate-co-butylene furandicarboxylate) copolyesters. Polymer Degradation and Stability 2013, 98 (11), 2177-2183. 26. Wang, G.; Jiang, M.; Zhang, Q.; Wang, R.; Tong, X.; Xue, S.; Zhou, G., Biobased copolyesters: Synthesis, sequence distribution, crystal structure, thermal and mechanical properties of poly(butylene sebacate-co-butylene furandicarboxylate). Polymer Degradation and Stability 2017, 143, 1-8. 27. Tokiwa, Y.; Calabia, B. P., Biodegradability and biodegradation of polyesters. Journal Of Polymers And the Environment 2007, 15 (4), 259-267.

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 30

28. Zorba, T.; Chrissafis, K.; Paraskevopoulos, K. M.; Bikiaris, D. N., Synthesis, characterization and thermal degradation mechanism of three poly(alkylene adipate)s: Comparative study. Polymer Degradation and Stability 2007, 92 (2), 222-230. 29. Rizzarelli, P.; Puglisi, C.; Montaudo, G., Soil burial and enzymatic degradation in solution of aliphatic co-polyesters. Polymer Degradation And Stability 2004, 85 (2), 855-863. 30. Bikiaris, D. N.; Achilias, D. S., Synthesis of poly(alkylene succinate) biodegradable polyesters 1. Mathematical modelling of the esterification reaction. Polymer 2006, 47 (13), 4851-4860. 31. Bikiaris, D. N.; Achilias, D. S., Synthesis of poly(alkylene succinate) biodegradable polyesters, Part II: Mathematical modelling of the polycondensation reaction. Polymer 2008, 49 (17), 3677-3685. 32. Lu, J.; Wu, L.; Li, B.-G., High Molecular Weight Polyesters Derived from Biobased 1,5-Pentanediol

and

a

Variety

of

Aliphatic

Diacids:

Synthesis,

Characterization,

and

Thermo-Mechanical Properties. Acs Sustainable Chemistry & Engineering 2017, 5 (7), 6159-6166. 33. Gigli, M.; Fabbri, M.; Lotti, N.; Gamberini, R.; Rimini, B.; Munari, A., Poly(butylene succinate)-based polyesters for biomedical applications: A review. European Polymer Journal 2016, 75, 431-460. 34. Gigli, M.; Negroni, A.; Soccio, M.; Zanaroli, G.; Lotti, N.; Fava, F.; Munari, A., Enzymatic hydrolysis studies on novel eco-friendly aliphatic thiocopolyesters. Polymer Degradation and Stability 2013, 98 (5), 934-942. 35. Siracusa, V.; Lotti, N.; Munari, A.; Rosa, M. D., Poly(butylene succinate) and poly(butylene succinate-co-adipate) for food packaging applications: Gas barrier properties after stressed treatments. Polymer Degradation and Stability 2015, 119, 35-45. 36. Park, J. H.; Jeon, J. Y.; Lee, J. J.; Jang, Y.; Varghese, J. K.; Lee, B. Y., Preparation of High-Molecular-Weight Aliphatic Polycarbonates by Condensation Polymerization of Diols and Dimethyl Carbonate. Macromolecules 2013, 46 (9), 3301-3308. 37. Jeon, J. Y.; Hwang, E. Y.; Eo, S. C.; Lee, B. Y., Preparation of macrodiols and polyols by chopping high-molecular-weight aliphatic polycarbonates. Journal of Polymer Science Part A: Polymer Chemistry 2014, 52 (11), 1570-1580. 38. Zhu, W.; Huang, X.; Li, C.; Xiao, Y.; Zhang, D.; Guan, G., High-molecular-weight aliphatic polycarbonates by melt polycondensation of dimethyl carbonate and aliphatic diols: synthesis and characterization. Polymer International 2011, 60 (7), 1060-1067. 39. Liu, W.; Zhu, W.; Li, C.; Guan, G.; Zhang, D.; Xiao, Y.; Zheng, L., Thermal degradation mechanism of poly(hexamethylene carbonate). Polymer Degradation and Stability 2015, 112, 70-77. 40. Zhu, W.; Li, C.; Zhang, D.; Guan, G.; Xiao, Y.; Zheng, L., Thermal degradation mechanism of poly(butylene carbonate). Polymer Degradation and Stability 2012, 97 (9), 1589-1595. 41. Wang, J.; Zheng, L.; Li, C.; Zhu, W.; Zhang, D.; Xiao, Y.; Guan, G., Fully biodegradable blends of poly(butylene succinate) and poly(butylene carbonate): Miscibility, thermal properties, crystallization behavior and mechanical properties. Polymer Testing 2012, 31 (1), 39-45. 42. Papageorgiou, G. Z.; Vassiliou, A. A.; Karavelidis, V. D.; Koumbis, A.; Bikiaris, D. N., Novel poly(propylene terephthalate-co-suceinate) random copolymers: Synthesis, solid structure, and enzymatic degradation study. Macromolecules 2008, 41 (5), 1675-1684. 43. Terzopoulou, Z.; Tsanaktsis, V.; Bikiaris, D. N.; Exarhopoulos, S.; Papageorgiou, D. G.; Papageorgiou, G. Z., Biobased poly(ethylene furanoate-co-ethylene succinate) copolyesters: solid state structure, melting point depression and biodegradability. Rsc Advances 2016, 6 (87), 84003-84015.

ACS Paragon Plus Environment

Page 29 of 30 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

44. Wang, G.; Jiang, M.; Zhang, Q.; Wang, R.; Zhou, G., Biobased copolyesters: synthesis, crystallization

behavior,

thermal

and

mechanical

properties

of

poly(ethylene

glycol

sebacate-co-ethylene glycol 2,5-furan dicarboxylate). Rsc Advances 2017, 7 (23), 13798-13807. 45. Soccio, M.; Costa, M.; Lotti, N.; Gazzano, M.; Siracusa, V.; Salatelli, E.; Manaresi, P.; Munari, A., Novel fully biobased poly(butylene 2,5-furanoate/diglycolate) copolymers containing ether linkages: Structure-property relationships. European Polymer Journal 2016, 81, 397-412. 46. Kwiatkowska, M.; Kowalczyk, I.; Kwiatkowski, K.; Szymczyk, A.; Jędrzejewski, R., Synthesis and structure – property relationship of biobased poly(butylene 2,5-furanoate) – block – (dimerized fatty acid) copolymers. Polymer 2017, 130, 26-38. 47. Guidotti, G.; Gigli, M.; Soccio, M.; Lotti, N.; Gazzano, M.; Siracusa, V.; Munari, A., Poly(butylene 2,5-thiophenedicarboxylate): An Added Value to the Class of High Gas Barrier Biopolyesters. Polymers 2018, 10 (2), 167. 48. Long, Y.; Zhang, R.; Huang, J.; Wang, J.; Jiang, Y.; Hu, G.-h.; Yang, J.; Zhu, J., Tensile Property Balanced and Gas Barrier Improved Poly(lactic acid) by Blending with Biobased Poly(butylene 2,5-furan dicarboxylate). ACS Sustainable Chemistry & Engineering 2017, 5 (10), 9244-9253. 49. Burgess, S. K.; Kriegel, R. M.; Koros, W. J., Carbon Dioxide Sorption and Transport in Amorphous Poly(ethylene furanoate). Macromolecules 2015, 48 (7), 2184-2193. 50. 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 (18), 4748-4756. 51. Sousa, A. F.; Matos, M.; Freire, C. S. R.; Silvestre, A. J. D.; Coelho, J. F. J., New copolyesters derived from terephthalic and 2,5-furandicarboxylic acids: A step forward in the development of biobased polyesters. Polymer 2013, 54 (2), 513-519.

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

For Table of Contents Use Only

Bio-based

biodegradable

poly(butylene

carbonate-co-furandicarboxylate)

copolyesters has potential application in plastic industry.

ACS Paragon Plus Environment

Page 30 of 30