Synthesis, physical properties, and biodegradability of Bio-based Poly

Each polymer was compressed to a 20 mm × 20 mm × 0.1 mm film at ..... WBS(Tg-TgBS) + k·WBO(Tg-TgBO) = 0. (3). 100. 80. 60. 40. 20. 0. W e ig h t/ %...
2 downloads 0 Views 1MB Size
Subscriber access provided by BOSTON UNIV

Article

Synthesis, physical properties, and biodegradability of Biobased Poly(butylene succinate-co-butylene oxabicyclate) Yuya Tachibana, Masayuki Yamahata, Saori Kimura, and Ken-ichi Kasuya ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/acssuschemeng.8b02112 • Publication Date (Web): 02 Jul 2018 Downloaded from http://pubs.acs.org on July 7, 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 35 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, physical properties, and biodegradability of Bio-based Poly(butylene succinate-co-butylene oxabicyclate) Yuya Tachibana1,2, Masayuki Yamahata1, Saori Kimura1, and Ken-ichi Kasuya1,2* 1 Division

of Molecular Science, Faculty of Science and Technology, Gunma University, 1-5-1

Tenjin, Kiryu, Gunma, 376-8515, Japan 2 Gunma

University Center for Food Science and Wellness, 4-2 Aramaki, Maebashi, Gunma,

371-8510, Japan

e-mail: [email protected]

ACS Paragon Plus Environment

1

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 2 of 35

ABSTRACT:

Furfural, a compound easily manufactured from hemicellulose on the global scale, has gained attention as a second-generation biomass resource. The synthesis of oxabicyclodicarboxylic anhydride (OBCA) from furfural alone and its polymerization with diol to form the bio-based and biodegradable polyoxabicyclate (POBC) make OBCA a novel bio-based monomer building block. To control the physical properties of POBC, we synthesized a copolyester of butanediol (BD), succinic anhydride (SAh), and OBCA. The structures of the resultant poly(butylene succinate-cobutylene oxabicyclate) (P(BS-co-BO)) polymers were studied using

1H

and

13C

NMR

spectrometry, size exclusion chromatography, thermal gravimetric analysis, differential scanning calorimetry, and wide-angle X-ray diffraction analysis. Physical properties of the melt-pressed films were examined by tensile strength testing, and the hydrolyzability of P(BS-co-BO)s was evaluated using a clear zone formation method with bacteria. To examine the environmental biodegradability of P(BS-co-BO), the biochemical oxygen demand for degrading their constituents was measured using the mixed inoculum method.

KEYWORDS:

Furfural,

Bio-based

Polymer,

Poly(butylene

succinate),

Copolyester,

Environmental biodegradabllity

ACS Paragon Plus Environment

2

Page 3 of 35 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 New bio-based building blocks and the associated bio-based polymers help to replace fossil fuels

with biomass to produce important chemicals on a commercial scale.1–5 Several companies uses bio-based compounds, namely lactic acid,6 succinic acid (SA),7 and propanediol8 as monomers to produce poly(lactic acid),9 poly(butylene succinate) (PBS),7 and poly(propylene terephthalate),10 respectively. Recently, some compounds produced from sugar have been added to the bio-based building block family. Isosorbide, isomannide, and isoidide were used as monomers of bio-based polycarbonate that was an alternative to the bisphenol-A-type polycarbonate.11,12 Ciclyc acetalized sugars were also used as monomers of bio-based polyester.13,14 Furandicarboxylic acid was used as a monomer of poly(ethylene furanoate) (PEF), a substitute of poly(ethylene terephthalate).15,16 These novel bio-based building blocks promoted the development of bio-based polymers including their copolymers.17–20 Until the early 21st century, bio-based building blocks have been mainly produced from edible biomass resources (i.e., first-generation biomass) like sugar, starch, and vegetable oil because of the effective fermentation process. Recently, inedible biomass resources, called second-generation biomass, such as lignocellulose (comprising of lignin, cellulose, and hemicellulose) provide a new approach that does not compete with food production.21 Furfural is manufactured from hemicellulose worldwide in large quantities. As one of the most value-added chemicals derived from biomass according to the U.S. Department of Energy, it has gained attention as a second-generation-biomass resource.22–25 The development of furfural as a chemical resource was actively pursued until middle 20th century, when progresses in the petroleum industry put furfural at a disadvantage due to the cost effectiveness, at a time before the importance of sustainability was acknowledged. As a result, the research of furfural as a chemical

ACS Paragon Plus Environment

3

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 4 of 35

resource declined, and its main industrial applications today are furan resin and extraction solvent.22 While Gandini et al. used furfural as a resource for polymers,5 our research group had reported a range of bio-based building blocks prepared from furfural alone for polymer synthesis, including SA,26 1,4-butanediol (BD),26,27 and terephthalic acid28 which were monomers of commercially available plastics such as poly(butylene succinate), poly(butylene terephthalate), and poly(ethylene terephthalate). We also synthesized oxabicyclodicarboxylic anhydride (OBCA) from furfural alone as a novel bio-based building block, and polymerized it with several diols to form polyoxabicyclates (POBCs).29 The resultant POBCs were amorphous polyesters with transparency and flexibility, because the bulky OBCA unit could not form crystalline phase. Furthermore, the POBCs have environmental biodegradability, which is a valuable property from the perspective of sustainability.1,30 POBCs can potentially replace general-purpose plastics, even though their physical properties like strength and thermal properties are still lacking on various occasions. The mechanical and thermal properties and biodegradability in polymers are significantly related to the crystallinity; and copolymerization with different units can control the crystalline phase. PBS as a crystalline and biodegradable polyester has been industrially produced from biomass. However, it lacks flexibility, and so copolymerizing it with a unit that controls the crystallinity may be desirable. For example, the physical properties of poly(butylene succinate-co-butylene adipate) (PBSA, the copolymer of PBS with poly(butylene adipate) (PBA), which disturbs the crystallinity of PBS) depend on the ratio of PBS and PBA units. The copolymerization with cyclic acetalized sugars, which depressed the crystallizability of PBS, provided flexibility and higher glass transition temperature.13,14 The copolymerization with PEF as a rigid unit provided further

ACS Paragon Plus Environment

4

Page 5 of 35 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

strength to PBS.31,32 Since BD, SA, and OBCA are potentially bio-based building blocks produced from furfural alone, and poly(butylene oxabicyclate) (PBO) produced from BD and OBCA has the best physical properties and biodegradability among POBCs,30 we designed a copolymer between PBS and PBO as a new bio-based material. PBS is a crystalline polymer and PBO is an amorphous polymer composed of a bulky oxabicyclate unit, so the introduction of butylene oxabicyclate (BO) unit to butylene succinate (BS) unit would disturb the crystalline phase of PBS and provide flexibility to PBS. In this study, we synthesized poly(butylene succinate-co-butylene oxabicyclate) (P(BS-co-BO)) polymers in a full range of compositions by direct esterification with titanium catalyst. The P(BSco-BO)s were characterized by 1H and

13C

NMR spectrometry, size exclusion chromatography

(SEC), thermal gravimetric analysis (TGA), differential scanning calorimetry (DSC), and wideangle X-ray diffraction analysis (WAXD). Their physical properties were examined by tensile strength testing using melt-pressed films. We also examined their biodegradability by microorganisms that could degrade polyesters using a clear zone formation method. Their environmental biodegradability was studied based on biochemical oxygen demand (BOD) biodegradation test and analysis of the hydrolysis products using the mixture inoculum method.



EXPERIMENTAL SECTION Chemicals and materials. BD was purchased from Tokyo Kasei Industries Co., Ltd. (Tokyo,

Japan) and used after distillation under reduced pressure. Titanium tetraisopropoxide was purchased from Wako Pure Chemical Industries, Ltd. (Osaka, Japan) and used without further purification. Succinic anhydride (SAh) and deuterated chloroform were purchased from Kanto

ACS Paragon Plus Environment

5

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 6 of 35

Chemical Industry Co., Ltd. (Tokyo, Japan) and used without further purification. OBCA was prepared following our previous study.29 General procedure of copolymerization. Given amounts of OBCA and SAh together with an excess amount of BD (2.1 equimolar of the sum of OBCA and SAh) were added to a roundbottom flask. The reaction mixture was heated to 160 °C for 3 h under dry N 2 flow to remove BD, which is a by-product of esterification. After cooling to room temperature, titanium tetraisopropoxide (0.34 mol% for acid anhydrides) catalyst was added to the mixture. Polycondensation was carried out under 10 Pa at 180 °C for 7 h. The resulting solid was dissolved in chloroform, and the solution was poured into methanol. The precipitate was filtered and dried under vacuum at 80 °C to give a white solid. 1

H and

13

C NMR characterization. To investigate the chemical structures and the

composition ratios of P(BS-co-BO)s, 1H and 13C NMR spectra were recorded on a 600 MHz NMR spectrometer (JNM-ECA600; JEOL, Tokyo, Japan) using deuterated chloroform. Measurement of molecular weight. Molecular weights were determined by SEC with a refractive index detector, using a combination of two TSKgel MultiporeHXL-M columns and a guard column (TSKgel guard column MP (XL)). All columns were from Tosoh Co., Tokyo, Japan. The columns were eluted with chloroform (1.0 mL min-1 at 40 °C) and polystyrene standards were used for calibration. Evaluation of thermal properties. Thermal properties were determined using DSC (SSC/5520; Seiko Instruments Inc., Chiba, Japan). After the sample was heated to 150 °C and cooled to −60 °C at a rate of 10 °C/min, it was heated to 150 °C at the same rate during the second heating scan. The glass transition temperature (Tg) and melting temperature (Tm) of each polymer were determined from the second heating scan, and the crystallization temperature (Tc) was

ACS Paragon Plus Environment

6

Page 7 of 35 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

determined from the first cooling scan. The curves of Tg and Tm as a function of the composition ratio of BO unit were fitted using Igor Pro ver. 6.37 (Wavemetrics, Ohio, USA) software. Melt-press molding. Each polymer was compressed to a 20 mm × 20 mm × 0.1 mm film at 15 MPa using a hot-pressing machine (Mini Press Test 10; Toyo Seiki Seisaku-sho Ltd., Tokyo, Japan) at 110 °C for 1 min. After compressing, the melt-pressed film was immediately cooled to room temperature using ice water. Crystallinity measurement. The WAXD measurements were conducted using an X-ray diffractometer (RINT220; Rigaku Co., Tokyo, Japan) to determine the degree of crystallinity (Xc) using Cu-Kα radiation with a wavelength of 0.154 nm. The voltage was set to 40 kV and the current 20 mA. Each melt-pressed film was mounted on a sample holder and scanned from 5˚ to 120˚. The Xc value was estimated according to the following equation: Xc = [Ic/(Ic + Ia)] × 100

(1)

where Ic and Ia are the scattering intensities of the crystalline peaks and amorphous halo, respectively. Tensile strength testing. Each melt-pressed film was cut into strips (20 mm × 2.0 mm × 0.1 mm). The strength and strain of the specimens at the breaking point were measured by tensile strength tests on a universal material testing machine (EZ-test; Shimadzu) at 23 °C. The grip distance was 10 mm, and the tensile strength test speed was 10 mm/min. The tensile strength was determined as the maximum strength required to break the materials on the stress-strain curve. The tensile strain at the breaking point was determined by the maximum strain on the stress-strain curve. The average value is taken over five samples measured under the same conditions. Preparation of emulsified media containing P(BS-co-BO).33 The P(BS-co-BO) polymers were dissolved in dichloromethane (30 mL). The solution was emulsified with an

ACS Paragon Plus Environment

7

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 35

ultrasonic disruptor (UD-200; TOMY Seiko Co. Ltd., Tokyo, Japan) in a basal medium (1 L) at pH = 7.0 with the following components (in wt./vol%): KH2PO4, 0.46; Na2HPO4·12H2O, 1.16; MgSO4·7H2O, 0.05; FeCl2·6H2O, 0.01; NH4Cl, 0.1; yeast extract, 0.01; and Plysurf, 0.01. Dichloromethane was removed by heating at 40 °C with a heating magnetic stirrer. Agar (15 g) was added to the emulsified medium, and the mixture was autoclaved at 110 °C for 20 min. The prepared P(BS-co-BO) media were solidified in a petri dish after cooling. Hydrolysis of polymer substrates by different bacterial strains.34,35 A total of 6 bacterial strains (NKCM2511,36 NKCM2512,37 MKCM1001, MKCM4002, TBKT040311B, and JKCM G-7A38) were tested for the degradation of ten P(BS-co-BO) polymers, PBAT, PCL, and PHB by the clear zone formation method. The accession numbers of 16S rDNA for the 6 strains are AB591807, AB591806, AB627010, AB691777, LC034566, and LC010671, respectively. BOD biodegradation test.33 The BOD biodegradability of P(BS-co-BO)s was determined by measuring oxygen consumption with a BOD tester (BOD tester 200F with a 300-mL BOD reactor, TAITEC Co., Koshigaya, Japan) by ISO 14851 standard at 25 °C. The BOD medium contained 200 mL of oxygen-saturated water and 0.2 mL of mineral salts from a stock solution with the following composition (in g·L−1): KH2PO4, 8.5; K2HPO4, 1.75; Na2HPO4·H2O, 33.3; NH4Cl, 1.7; MgSO4·7H2O, 8.5; CaCl2, 27.5; and FeCl3·6H2O, 0.25. A P(BS-co-BO) sample (about 10 mg) was placed in the 300-mL BOD reactor, and 200 mL of minimal medium was added to the reactor. The mixture inoculum (200 μL) was added to a BOD reactor, and the reactor was incubated at 25 °C. The BOD biodegradability was defined by following equation: BOD biodegradability (%) = (BOD sample – BODblank)/ThOD × 100

(2)

ACS Paragon Plus Environment

8

Page 9 of 35 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

where BOD sample and BODblank are the experimental values of oxygen demand of the sample and a blank medium, respectively. ThOD is the theoretical oxygen demand of the sample calculated by assuming that the film was completely degraded into CO2 and H2O.



RESULTS AND DISCUSSION Copolymerization and molecular structure of P(BS-co-BO)s. The polymerization of

OBCA, BD, and SAh was carried out as shown in Scheme 1. The two-step polycondensation gave P(BS-co-BO) with a given composition ratio in the BS and BO units (denoted as P(BS(100-x)-coBOx), x = 0–100). The data for the homopolymer P(BS0-co-BO100) were obtained in the previous reports29,30 and reprinted here.

Scheme 1.

The feed ratio of SAh and OBCA, the composition ratio of butylene succinate (BS) and butylene oxabicyclate (BO) units, the yield, and the molecular weights of P(BS-co-BO)s are summarized in Table 1. Different feed ratios of SAh and OBCA were used to produce 10 P(BS-co-BO)

ACS Paragon Plus Environment

9

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

copolymers. Figures 1a and 1b show the partial 1H and

13C

Page 10 of 35

NMR spectra (for full spectra see

Figure S1 and S2), which reveal the composition ratio and the sequence distribution. The peaks were assigned according to the previous study.29 The composition ratio was determined by the integrated peak area ratio between H f of carbonyl α-methine protons in the BO unit at 2.98 ppm and Hc of carbonyl α-methyl protons in the BS unit at 2.63 ppm. The composition ratio of BO unit in the copolymer is smaller than the feed ratio of OBCA, indicating that OBCA has lower reactivity than SAh due to the bulkiness of its oxabicyclo unit. As OBCA was considerably sublimated during the polycondensation under reduced pressure, the yields of all P(BS-co-BO) polymers were low. The low reactivity of OBCA compared to its sublimation further lowered the yield of P(BSco-BO), in which the OBCA unit ratio was lower than the SA unit ratio. The molecular weight of all copolymers was more than 1 × 104 and did not depend on the composition ratio.

ACS Paragon Plus Environment

10

Page 11 of 35 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

ACS Sustainable Chemistry & Engineering

Table 1. Molar composition, yield, and molecular weights of P(BS-co-BO)sa –––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– Molar composition P(BS-co-BO)

–––––––––––––––––––––––––––––––––––––––––– Feed ratio Composition ratiob

Yield c/ %

Mnd × 10−4

Mw/Mnd

––––––––––––––––––––– –––––––––––––––––––– SAh OBCA XBS XBO –––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– P(BS100-co-BO0) 1.00 0.00 1.00 0.00 68 3.11 1.75 P(BS99-co-BO1)

0.95

0.05

0.99

0.01

73

4.01

1.66

P(BS97-co-BO3)

0.90

0.10

0.97

0.03

53

4.09

2.06

P(BS90-co-BO10)

0.75

0.25

0.90

0.10

42

6.04

2.13

P(BS70-co-BO30)

0.50

0.50

0.70

0.30

38

5.50

1.74

P(BS66-co- O34)

0.25

0.75

0.66

0.34

45

2.28

1.42

P(BS56-co-BO44)

0.10

0.90

0.56

0.44

43

3.07

2.56

P(BS13-co-BO87)

0.05

0.95

0.13

0.87

31

3.78

3.67

P(BS7-co-BO93)

0.01

0.99

0.07

0.93

27

1.44

3.30

P(BS0-co-BO100)

0.00

1.00

0.00

1.00

35

2.32

1.38

–––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– a Reaction condition: 160 °C under atmospheric pressure for 3 h, and then at 180 °C under 10 Pa for 7 h. b Calculated by 1H NMR. c Methanol-insoluble parts. d Estimated by SEC (eluent: chloroform) calibrated with polystyrene standards.

ACS Paragon Plus Environment

11

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 12 of 35

Figure 1. a) 1H NMR (600 MHz) spectra and partial 13C NMR (151 MHz) spectra at b) 173–171 ppm and c) 63.8–64.6 ppm for the P(BS-co-BO) copolymers. c) Possible triads of P(BS-co-BO)s. d) Three triads in P(BS-co-BO)s.

The sequence analysis of units is important in copolymers, where the physical properties are influenced by the sequence distribution.39 In the case of copolyester, the integrated peak ratio of methylene carbons in 13C NMR spectra give the sequence distribution.40,41 Figure 1d displays three triads assumed as the sequence of P(BS-co-BO), and Figure 1b displays partial 13C NMR spectra

ACS Paragon Plus Environment

12

Page 13 of 35 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 methylene carbon attached to the oxygen atom. The peak of CO1 in oxabicyclate-butyleneoxabicyclate (OBO) triad, that of CS1 in succinate-butylene-succinate (SBS) triad, and that of CO2 and CS2 in succinate-butylene-oxabicyclate (SBO) triad were observed at 64.35, 64.04, 64.26, and 64.15 ppm, respectively. The integrated peak ratio (CO1 + CO2 / CS1 + CS2) in 13C NMR spectra of P(BS-co-BO)s was almost identical to the composition ratio calculated from 1H NMR spectra. The CS1 peak was not observed, but CO1, CO2, and CS2 peaks were found in the spectra of P(BS7-coBO93) and P(BS13-co-BO87), indicating that the SBS triad is not in them. The CO1 peak was not observed for P(BS99-co-BO1), P(BS97-co-BO3), P(BS90-co-BO10), P(BS70-co-BO30), P(BS66-co-BO34), and P(BS56-co-BO44), indicating that the OBO triad is not in them. However, this result conflicts with the integrated peak ratio between CO2 and CS2, which should be equal. In addition, the peak at 64.27 ppm derived from the irregular structure of P(BS-co-BO)29 disturbed the sequence analysis because it overlapped with the peak of CO2. We assume that all P(BS-coBO)s are random copolymers, and the peaks of CO2 and CS2 in SBO triad would coalesce with those of CO1 in OBO and CS2 in SBS triad, respectively. The CS1 peaks of P(BS70-co-BO30), P(BS66-co-BO34), and P(BS56-co-BO44) shifted slightly up-field, supporting our assumption.

Thermal properties. The thermal stability of P(BS-co-BO)s was evaluated using TGA. Figure 2 displays the TGA curves. The 5% weight loss temperatures are summarized in Table 2, being all above 290 °C except for P(BS13-co-BO87) and P(BS7-co-BO93). The broad polydispersity of P(BS13-co-BO87) and P(BS7-co-BO93) shown in Table 1 indicates that they include lowmolecular-weight fractions that were thermally less stable than high-molecular-weight fractions. Therefore, the thermal stability of P(BS13-co-BO87) and P(BS7-co-BO93) became lower than the others.

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering

100

80

– – – P(BS100-co-BO0) – – – P(BS99-co-BO1)

Weight/ %

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 35

60

– – – P(BS97-co-BO3) – – – P(BS90-co-BO10) –––– P(BS70-co-BO30)

40

–––– P(BS66-co-BO34) – – – P(BS56-co-BO44) –––– P(BS13-co-BO87)

20

– – – P(BS7-co-BO93) –––– P(BS0-co-BO100)

0 150

200

250 300 350 Temperature/ °C

400

450

Figure 2. TGA curves of P(BS-co-BO)s obtained using a temperature ramp rate of 10 °C·min-1.

The Tg, Tm, and Tc values of P(BS-co-BO)s were determined using DSC. Figure 3a and b display the DSC charts of P(BS-co-BO)s during the first cooling and the second heating. Tm, Tc, the enthalpy of fusion (ΔHm), and the enthalpy of crystallization (ΔHc) are summarized in Table 2. The Tg of P(BS99-co-BO1), P(BS97-co-BO3), and P(BS90-co-BO10) could not be observed in DSC measurement due to their high crystallinity. P(BS70-co-BO30), P(BS66-co-BO34), P(BS56co-BO44), P(BS13-co-BO87), P(BS7-co-BO93), and P(BS0-co-BO100) showed a single Tg that increased from -24 °C to 32 °C with increasing composition ratio of the BO unit, indicating that no phase separation between BS and BO units took place in the amorphous phase. P(BS100-coBO0) obtained in this study did not show the glass transition, while the T g of PBS is generally observed at around -36 °C.42 The relationship between Tg values and the composition ratio of random copolymers can be estimated according to Wood’s equation as43,44 WBS(Tg-TgBS) + k·WBO(Tg-TgBO) = 0

(3)

ACS Paragon Plus Environment

14

Page 15 of 35 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

where Tg (K) is the glass transition temperature of the copolymer; TgBS (K) and TgBO (K) are the Tg of P(BS100-co-BO0) and P(BS0-co-BO100); WBS and WBO are the weight fractions of BS unit and BO units, respectively; and k is a constant characteristic of P(BS-co-BO). As TgBO > TgBS, equation 3 can be rearranged as Tg = [TgBS + (k·TgBO – TgBS) WBO]/[1-(1- k) WBO]

(4)

Figure 3c plots Tg as a function of WBO following equation 4. Wood’s equation fitted the experimental data well with k = 0.504, indicating that P(BS-co-BO)s are random copolymers. The values of Tm, Tc, ΔΗ m, and ΔΗc are plotted as a function of the composition ratio of BO unit in Figure S2. Those of P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), and P(BS90-coBO10) decreased with increasing BO unit ratio, indicating that these units disturbs the formation of crystalline phase of BS unit. The relationship between T m and the composition ratio of random copolymer can be estimated according to Flory’s equation44,45 as 1/Tm – 1/TmBS = – (R/ΔHBS) lnXBS

(5)

where TmBS is the melting point of P(BS100-co-BO0), R is gas constant, HBS is the heat of fusion per BS unit, and XBS is the molar composition ratio of BS unit. Figure 3d displays a plot of 1/Tm as a function of lnXBS. The value of HBS obtained from the slope of the straight line is 7.8 kJ/mol. Flory’s equation fitted well the experimental data, also indicating that P(BS-co-BO)s are random copolymers. In addition, the Tm and Tc of P(BS-co-BO)s depend on the composition ratio of BO unit, indicating that these polymers have isomorphic crystalline structures.

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering

a)

b)

320

c)

2.7

d)

Exo.

P(BS100-co-BO0)

300

P(BS99-co-BO1) P(BS97-co-BO3) P(BS90-co-BO10)

P(BS66-co-BO34)

-3

Tg / K

1/Tm x10 / K

-1

280

P(BS70-co-BO30)

Heat flow

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 16 of 35

260

2.6

P(BS56-co-BO44) P(BS13-co-B87)

240 P(BS7-co-BO93) P(BS0-co-BO100)

220

-50

0 50 100 Temperature/ °C

-50

0 50 100 Temperature/ °C

2.5 0.0

0.5 WBO

1.0

0.00

0.05 -lnXBS

0.10

Figure 3. DSC charts of P(BS-co-BO)s in a) the first cooling cycle and b) the second heating cycle. Cooling and heating rates: 10 °C·min-1. c) Tg versus the weight fraction of BO unit (WBO). The curve represents the theoretical value generated from Eq. 4 with k = 0.504. d) 1/Tm versus the natural logarithm of BS unit (-lnXBS). The theoretical plot generated from Eq. 5 is represented as a straight line with slope.

ACS Paragon Plus Environment

16

Page 17 of 35 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 2. Thermal properties and crystallinity of P(BS-co-BO)s ––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– P(BS-co-BO) Td5%a Tgb Tmb Tcb ΔHmb ΔHcb Xcc / °C / °C / °C / °C / J·g-1 / J·g-1 /% ––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– P(BS100-co-BO0)

310.2

-d

116.2

78.1

83

72

74

P(BS99-co-BO1)

307.1

-d

115.4

71.7

71

70

63

P(BS97-co-BO3)

302.6

-d

111.5

70.2

66

63

51

P(BS90-co-BO10)

306.8

-d

100.2

50.6

46

45

41

P(BS70-co-BO30)

304.2

-24.2

-

-

-

-

17

P(BS66-co-BO34)

334.1

-20.3

-

-

-

-

-

P(BS56-co-BO44)

305.0

-8.8

-

-

-

-

-

P(BS13-co-BO87)

240.6

20.0

-

-

-

-

-

P(BS7-co-BO93)

293.5

28.1

-

-

-

-

-

P(BS0-co-BO100)

290.1

31.6

-

-

-

-

-

––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––––– a Measured by thermal gravimetric analysis (TGA). b Measured by differential scanning calorimetry (DSC).

c

Calculated using X-ray diffractometer (WAXD).

d

Not observed in DSC

measurement.

Crystallinity. The WAXD patterns were measured using molded self-standing films after melt pressing the co-polymers P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), P(BS90-coBO10), P(BS70-co-BO30), and P(BS0-co-BO100). In contrast, P(BS70-co-BO30), P(BS66-coBO34), P(BS56-co-BO44), P(BS13-co-BO87), and P(BS7-co-BO93), of which Tg was below 28 °C and a melting peak was not observed, were not evaluated. In the WAXD patterns in Figure 4a, the diffraction peaks from [020] and [110] planes of PBS46 were observed at around 19.6° and 22.6° in P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), P(BS90-co-BO10), and P(BS70-

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering

co-BO30). Meanwhile, only a halo peak was observed in P(BS0-co-BO100), whose melting point was not observed. The 2θ values of [020] and [110] planes slightly decreased, and the change in d spacing as a function of the composition ratio of BO unit was plotted in Figure 4b. The crystallinity of P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), and P(BS90-co-BO10) also decreased with increasing ratio of BO unit as summarized in Table 2. These results suggest that the bulky oxabicyclate could disturb the crystalline phase of PBS, would expand the d spacing, and decrease the crystallinity.47 In addition, P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), P(BS70co-BO30), and P(BS90-co-BO10) are crystalline polyesters whose thermal properties depend on the composition ratio of BS unit. Although diffraction peaks of PBS were observed in the WAXD pattern of P(BS70-co-BO30), no exothermic and endothermic peaks derived from crystallization were observed in the DSC measurement because the crystallinity was low.

0.48

a)

b)

0.46

P(BS100-co-BO0)

d / nm

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 35

P(BS99-co-BO1)

0.44

P(BS97-co-BO3)

0.42 P(BS90-co-BO10)

0.40

P(BS70-co-BO30) P(BS0-co-BO100)

0.38 10

20 30 2q/ °

40

0 10 20 30 Composition ratio of BO unit/ %

Figure 4. a) WAXD patterns of selected P(BS-co-BO) films. b) Experimental diffraction spacing (d) of the PBS crystalline phase vs. the composition ratio of BO unit ratio. Closed triangle: [020] plane. Open circle: [110] plane.

ACS Paragon Plus Environment

18

Page 19 of 35

Mechanical properties. The mechanical properties of P(BS100-co-BO0), P(BS99-co-BO1), P(BS97-co-BO3), P(BS90-co-BO10), P(BS70-co-BO30), and P(BS0-co-BO100) were measured using tensile strength testing. Figure 5a displays the obtained stress-strain curves. The Young’s modulus, tensile strength, and strain at breaking point are plotted in Figure 5b. The Young’s modulus and tensile strength increased with increasing ratio of BS unit, while the strain at the breaking point decreased, indicating that the crystalline phase derived from BS unit improves the hardness of P(BS-co-BO)s. This result indicates that the bulky BO unit disturbs the crystallinity of BS unit and gives high flexibility to PBS, which is a hard plastic. On the other hand, as P(BS0co-BO100) is an amorphous polymer, its physical properties do not follow this tendency.

b)

a) 30

600

40

800

30

600

20

400

10

200

– P(BS100-co-BO0)

Stress/ MPa

– P(BS70-co-BO30)

15

– P(BS0-co-BO100)

10

~ ~ ~ ~

5 0 0

10

20

30 200

400

200

Tensile Strength/ MPa

– P(BS90-co-BO10)

20

Young's modulus/ MPa

– P(BS97-co-BO3)

0 400 Strain/ %

600

800

0

Strain at beaking point/ %

– P(BS99-co-BO1)

25

~~

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

0 0

20 40 60 80 Composition ratio of BS unit/ %

100

Figure 5. a) Stress-strain curves for P(BS-co-BO)s. b) Young’s modulus (red closed triangle), tensile strength (blue closed square), and strain at breaking point (black closed circle) of P(BS-coBO)s as a function of the composition ratio of BS unit.

Biodegradability of PBSOs. We previously reported that microorganisms readily mineralized all the hydrolysates of PBS48 and PBO,30 namely SA, BD, and oxabicyclic diacid.

ACS Paragon Plus Environment

19

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 35

Therefore, P(BS-co-BO)s may also be biodegradable and hydrolyzed by the extracellular enzymes produced by the microorganisms. These microorganisms can form a clear zone around their colonies by hydrolysis of P(BS-co-BO)s in emulsion, and the capacity to form such clear zones could be used to measure the enzymatic hydrolyzability of P(BS-co-BO)s. In this study, 6 bacterial strains that are known to degrade some commercial biodegradable polymers were inoculated on an emulsified medium containing P(BS-co-BO)s. For comparison, PBAT, PCL, PHB, and olive oil were also evaluated using the same 6 bacterial inocula. The results are summarized in Table 3. The boundary of a small clear zone is less than 1 mm from the colony, while those for medium and large clear zones are 1–5 and larger than 5 mm, respectively. •

On the emulsified medium containing P(BS100-co-BO0), P(BS99-co-BO1), P(BS70-coBO30), P(BS13-co-BO87), or P(BS7-co-BO93), none of the strains formed a clear zone.



For P(BS97-co-BO3) medium, only TBKT040311B related to Bacillus thuringnesis formed a small clear zone.



For P(BS90-co-BO10), two bacterial strains formed small clear zones: NKCM2512 belonging to the genus Rhodococcus and JKCM G-7A related to B. megaterium.



On P(BS66-co-BO34) and P(BS56-co-BO44) emulsified media, five strains formed large- or medium clear zones: NKCM2512, MKCM1001 belonging to the genus Variovorax, MKCM4002 related to B. altitudinus, TBKT04311B, and JKCM G-7A.

The PBAT and PCL-degrading enzyme is cutinase,36,49 the PHB-degrading enzyme is PHB depolymerase,50, and the olive oil-degrading enzyme is lipase. Although we could not find a relationship among the clear zone formation and the reported depolymerases for PBAT, PCL, PHB, and olive oil; several bacterial strains formed clear zones on the P(BS-co-BO) emulsified media. This indicates that some of these copolymers are biodegradable and could undergo

ACS Paragon Plus Environment

20

Page 21 of 35 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

hydrolyzation by microorganisms in natural environments. The enzymatic hydrolyzability of PBS depends on the crystallinity51 and that of copolyester depends on the length of homo sequence of each unit52. The high clear zone formation ability on P(BS66-co-BO34) and P(BS56-co-BO44) emulsified media, indicating that the enzymatic hydrolyzability of P(BS-co-BO) increased with decreasing the length of homo sequence of BS and BO units.

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering

Table 3. Clear zone formation by polyester-degrading bacteria on plates containing various polyesters and olive oil in emulsified medium

Strain

Genus

P(BS100-co-BO0)

P(BS99-co-BO1)

P(BS97-co- BO3)

P(BS90-co- BO10)

P(BS70-co- BO30)

P(BS66-co- BO34)

P(BS56-co-BO44)

P(BS13-co-BO87)

P(BS7-co-BO93)

P(BS0-co-BO100)

PBAT

PCL

PHB

Olive oil

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

Page 22 of 35

NKCM2511

Rhodococcus

-

-

-

-

-

-

-

-

-

+

+

++

-

-

NKCM2512

Rhodococcus

-

-

-

+

-

++

+++

-

-

+

+

++

-

++

MKCM1001

Variovorax

-

-

-

-

-

+++

++

-

-

-

-

-

-

-

MKCM4002

Bacillus

-

-

-

-

-

++

++

-

-

-

-

+

-

-

TBKT040311B

Bacillus

-

-

+

-

-

++

++

-

-

-

+

+++

+

-

JKCM G-7A

Bacillus

-

-

-

+

-

++

++

-

-

-

-

-

+

-

+++: Large clear zone (> 5 mm) formed. ++: Medium clear zone (1–5 mm) formed. +: Small clear zone (< 1 mm) formed. -: No clear zone.

ACS Paragon Plus Environment

22

Page 23 of 35

100

80

BOD biodegradability/ %

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



P(BS100-co-BO0)



P(BS99-co-BO1)



P(BS97-co-BO3)



P(BS90-co-BO10)

— P(BS70-co-BO30)

60

40



P(BS66-co-BO34)



P(BS56-co-BO44)



P(BS13-co-BO87)

▲ ▲

P(BS7-co-BO93) P(BS0-co-BO100)

20

0 0

10

20

30

40

50

60

Incubation/ day

Figure 6. BOD biodegradation curves of P(BS-co-BO)s.

The environmental biodegradability of P(BS-co-BO)s was further evaluated using BOD biodegradation testing at 25°C, which well demonstrates the biodegradation in natural environment. Figure 6 shows the BOD biodegradation curves of P(BS-co-BO)s. Of the 10 polymers, 5 showed no BOD biodegradability using the inocula: P(BS99-co-BO1), P(BS97-coBO3), P(BS90-co-BO10), P(BS70-co-BO30), and P(BS13-co-BO87). For P(BS0-co-BO100), the BOD biodegradation started immediately and continued until day 60. The curves of P(BS100-coBO0), P(BS66-co-BO34), P(BS56-co-BO44), and P(BS7-co-BO93) showed a lag untill day 7, 6, 10, and 20; and their biodegradation continued until day 60, 20, 22, and 38, respectively. After 60 days, the respective BOD biodegradabilities of P(BS100-co-BO0), P(BS66-co-BO34), P(BS56co-BO44), P(BS7-co-BO93), and P(BS0-co-BO100) reached 54%, 22%, 19%, 25%, and 81%. In addition to P(BS56-co-BO44) and P(BS66-co-BO34), which 5 of 6 strains tested degraded (they could form the clear zone around the colony on the plates), P(BS7-co-BO93) showed a certain

ACS Paragon Plus Environment

23

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 24 of 35

level of biodegradability. The BOD biodegradability of homopolymers; P(BS100-co-BO0) and P(BS0-co-BO100), were also high although both were never degraded by all strains tested. Here, the microbial mineralization degree of the polymers, an ultimate biodegradability is accounted for by the BOD biodegradability, while the enzymatic hydrolyzability, a primary biodegradability is done by the clear zone formation ability. This is why the BOD biodegradability is not in good agreement with the result of clear zone formation. Meanwhile, P(BS56-co-BO44) and P(BS66-coBO34) were degraded well by enzymes produced by microorganisms, and the hydrolysates were mineralized at high levels. This indicates that at least the incorporation of BO unit into the backbone of PBS can improve the primary biodegradability of its by the strain tested in the study.



CONCLUSION In this article, OBCA was used as a bio-based and biodegradable co-monomer to improve PBS

that is a commercially available bio-based polymer. The polycondensation of OBCA, SAh, and BD was performed with titanium isopropoxide catalyst under reduced pressure at high temperature to afford 10 different P(BS-co-BO) polymers (including two homopolymers). The P(BS-co-BO)s were expected to be random copolymers from their thermal properties and XRD patterns. The copolymers with more than 10% BO unit were crystalline and could be molded into self-standing films. The physical properties indicate that BO is a good copolymer unit for PBS to give flexibility. Additionally, bacterial strains could form clear zones in the emulsions of P(BS-co-BO)s with around a half of BO unit and 5 P(BS-co-BO)s showed BOD biodegradability, indicating that the introduction of BO unit to PBS improves the biodegradability. These results demonstrate that the bio-based BO unit could control several properties of polyesters by copolymerization. Therefore, OBCA was demonstrated to be a new promising member of the bio-based building block family.

ACS Paragon Plus Environment

24

Page 25 of 35 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. 1H and 13C NMR spectra. This material is available free of charge via

the Internet at ***. ◼

AUTHOR INFORMATION

Corresponding Author Tel: +81-27730-1487, e-mail: [email protected] ORCID Yuya Tachibana: 0000-0002-4567-7061 Ken-ichi Kasuya: 0000-0001-6818-6946 Notes The authors declare no competing financial interest. Author Contributions Y.T. and K.K designed this work and wrote the paper. Y.T., M.Y., and S.K. carried out the synthetic experiments and evaluated the properties and biodegradability. All the authors participated in the analysis and discussion of the results. ◼

ACKNOWLEDGMENT Y.T. is grateful for the financial support from JSPS KAKENHI Grant-in-Aid for Young

Scientists (B) Number 24710087 and the Precursory Research for Embryonic Science and Technology (PRESTO) program of the Japan Science and Technology Agency (JST) Grant Number JPMJPR13B7.

ACS Paragon Plus Environment

25

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

◼ (1)

Page 26 of 35

REFERENCES Schneiderman, D. K.; Hillmyer, M. A. 50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers. Macromolecules 2017, 50 (10), 3733–3749, DOI 10.1021/acs.macromol.7b00293.

(2)

Zhu, Y.; Romain, C.; Williams, C. K. Sustainable Polymers from Renewable Resources. Nature 2016, 540 (7633), 354–362, DOI 10.1038/nature21001.

(3)

Della Pina, C.; Falletta, E.; Rossi, M. A Green Approach to Chemical Building Blocks. The Case of 3-Hydroxypropanoic Acid. Green Chem. 2011, 13 (7), 1624–1632, DOI 10.1039/c1gc15052a.

(4)

Harmsen, P. F. H.; Hackmann, M. M.; Bos, H. L. Green Building Blocks for Bio-Based Plastics. Biofuels, Bioprod. Biorefining 2014, 8 (3), 306–324, DOI 10.1002/bbb.1468.

(5)

Gandini, A.; Lacerda, T. M.; Carvalho, A. J. F.; Trovatti, E. Progress of Polymers from Renewable Resources: Furans, Vegetable Oils, and Polysaccharides. Chem. Rev. 2016, 116 (3), 1637–1669, DOI 10.1021/acs.chemrev.5b00264.

(6)

Wang, Y.; Tashiro, Y.; Sonomoto, K. Fermentative Production of Lactic Acid from Renewable Materials: Recent Achievements, Prospects, and Limits. J. Biosci. Bioeng. 2015, 119 (1), 10–18, DOI 10.1016/j.jbiosc.2014.06.003.

(7)

Jansen, M. L. A.; van Gulik, W. M. Towards Large Scale Fermentative Production of Succinic Acid. Curr. Opin. Biotechnol. 2014, 30, 190–197, DOI 10.1016/j.copbio.2014.07.003.

ACS Paragon Plus Environment

26

Page 27 of 35 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

(8)

Saxena, R. K.; Anand, P.; Saran, S.; Isar, J. Microbial Production of 1,3-Propanediol: Recent Developments and Emerging Opportunities. Biotechnol. Adv. 2009, 27 (6), 895– 913, DOI 10.1016/j.biotechadv.2009.07.003.

(9)

Maharana, T.; Mohanty, B.; Negi, Y. S. Melt-Solid Polycondensation of Lactic Acid and Its Biodegradability. Prog. Polym. Sci. 2009, 34 (1), 99–124, DOI 10.1016/j.progpolymsci.2008.10.001.

(10)

Liu, H.; Xu, Y.; Zheng, Z.; Liu, D. 1,3-Propanediol and Its Copolymers: Research, Development and Industrialization. Biotechnol. J. 2010, 5 (11), 1137–1148, DOI 10.1002/biot.201000140.

(11)

De Jong, E.; Higson, A.; Walsh, P.; Wellisch, M. Product Developments in the Bio-Based Chemicals Arena. Biofuels, Bioprod. Biorefining 2012, 6 (6), 606–624, DOI 10.1002/bbb.1360.

(12)

Dussenne, C.; Delaunay, T.; Wiatz, V.; Wyart, H.; Suisse, I.; Sauthier, M. Synthesis of Isosorbide: An Overview of Challenging Reactions. Green Chem. 2017, 19 (22), 5332– 5344, DOI 10.1039/C7GC01912B.

(13)

Zakharova, E.; Alla, A.; Martínez de Ilarduya, A.; Muñoz-Guerra, S. Bio-Based PBS Copolyesters Derived from a Bicyclic d-Glucitol. RSC Adv. 2015, 5 (57), 46395–46404, DOI 10.1039/C5RA03844H.

(14)

Zakharova, E.; Lavilla, C.; Alla, A.; Martínez De Ilarduya, A.; Muñoz-Guerra, S. Modification of Properties of Poly(Butylene Succinate) by Copolymerization with

ACS Paragon Plus Environment

27

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 35

Tartaric Acid-Based Monomers. Eur. Polym. J. 2014, 61, 263–273, DOI 10.1016/j.eurpolymj.2014.09.024. (15)

Eerhart, A. J. J. E.; Huijgen, W. J. J.; Grisel, R. J. H.; van der Waal, J. C.; de Jong, E.; de Sousa Dias, A.; Faaij, A. P. C.; Patel, M. K. Fuels and Plastics from Lignocellulosic Biomass via the Furan Pathway; a Technical Analysis. RSC Adv. 2014, 4 (7), 3536–3549, DOI 10.1039/C3RA43512A.

(16)

Codou, A.; Moncel, M.; van Berkel, J. G.; Guigo, N.; Sbirrazzuoli, N. Glass Transition Dynamics and Cooperativity Length of Poly(Ethylene 2,5-Furandicarboxylate) Compared to Poly(Ethylene Terephthalate). Phys. Chem. Chem. Phys. 2016, 18 (25), 16647–16658, DOI 10.1039/C6CP01227B.

(17)

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 2,5Furandicarboxylic Acid: A Tribute to Furan Excellency. Polym. Chem. 2015, 6 (33), 5961–5983, DOI 10.1039/C5PY00686D.

(18)

Chernyshev, V. M.; Kravchenko, O. A.; Ananikov, V. P. Conversion of Plant Biomass to Furan Derivatives and Sustainable Access to the New Generation of Polymers, Functional Materials and Fuels. Russ. Chem. Rev. 2017, 86 (5), 357–387, DOI 10.1070/RCR4700.

(19)

Delidovich, I.; Hausoul, P. J. C.; Deng, L.; Pfützenreuter, R.; Rose, M.; Palkovits, R. Alternative Monomers Based on Lignocellulose and Their Use for Polymer Production. Chem. Rev. 2016, 116 (3), 1540–1599, DOI 10.1021/acs.chemrev.5b00354.

ACS Paragon Plus Environment

28

Page 29 of 35 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

(20)

Gomes, M.; Gandini, A.; Silvestre, A. J. D.; Reis, B. Synthesis and Characterization of Poly(2,5-Furan Dicarboxylate)s Based on a Variety of Diols. J. Polym. Sci. Part A Polym. Chem. 2011, 49 (17), 3759–3768, DOI 10.1002/pola.24812.

(21)

Axelsson, L.; Franzén, M.; Ostwald, M.; Berndes, G.; Lakshmi, G.; Ravindranath, N. H. Perspective: Jatropha Cultivation in Southern India: Assessing Farmers’ Experiences. Biofuels, Bioprod. Biorefining 2012, 6 (3), 246–256, DOI 10.1002/bbb.

(22)

Hoydonckx, H. E.; Van Rhijn, W. M.; Van Rhijn, W.; De Vos, D. E.; Jacobs, P. A. Furfural and Derivatives. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2007, DOI 10.1002/14356007.a12_119.pub2.

(23)

Gandini, A.; Coelho, D.; Gomes, M.; Reis, B.; Silvestre, A. Materials from Renewable Resources Based on Furan Monomers and Furan Chemistry: Work in Progress. J. Mater. Chem. 2009, 19 (45), 8656–8664, DOI 10.1039/b909377j.

(24)

Gandini, A. Furans as Offspring of Sugars and Polysaccharides and Progenitors of a Family of Remarkable Polymers: A Review of Recent Progress. Polym. Chem. 2010, 1 (3), 245–251, DOI 10.1039/B9PY00233B.

(25)

Kawamoto, H.; Saito, S.; Hatanaka, W.; Saka, S. Catalytic Pyrolysis of Cellulose in Sulfolane with Some Acidic Catalysts. J. Wood Sci. 2007, 53 (2), 127–133, DOI 10.1007/s10086-006-0835-y.

(26)

Tachibana, Y.; Masuda, T.; Funabashi, M.; Kunioka, M. Chemical Synthesis of Fully Biomass-Based Poly(Butylene Succinate) from Inedible-Biomass-Based Furfural and

ACS Paragon Plus Environment

29

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 30 of 35

Evaluation of Its Biomass Carbon Ratio. Biomacromolecules 2010, 11 (10), 2760–2765, DOI 10.1021/bm100820y. (27)

Kunioka, M.; Masuda, T.; Tachibana, Y.; Funabashi, M.; Oishi, A. Highly Selective Synthesis of Biomass-Based 1,4-Butanediol Monomer by Alcoholysis of 1,4Diacetoxybutane Derived from Furan. Polym. Degrad. Stab. 2014, 109, 393–397, DOI 10.1016/j.polymdegradstab.2014.05.011.

(28)

Tachibana, Y.; Kimura, S.; Kasuya, K. Synthesis and Verification of Biobased Terephthalic Acid from Furfural. Sci. Rep. 2015, 5, 8249, DOI 10.1038/srep08249.

(29)

Tachibana, Y.; Yamahata, M.; Kasuya, K. Synthesis and Characterization of a Renewable Polyester Containing Oxabicyclic Dicarboxylate Derived from Furfural. Green Chem. 2013, 15 (5), 1318–1325, DOI 10.1039/c3gc36938b.

(30)

Tachibana, Y.; Yamahata, M.; Ichihara, H.; Kasuya, K. Biodegradability of Polyesters Comprising a Bio-Based Monomer Derived from Furfural. Polym. Degrad. Stab. 2017, 146, 121–125, DOI 10.1016/j.polymdegradstab.2017.10.001.

(31)

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 Adv. 2016, 6 (87), 84003–84015, DOI 10.1039/C6RA15994J.

(32)

Morales-Huerta, J. C.; Ciulik, C. B.; de Ilarduya, A. M.; Muñoz-Guerra, S. Fully BioBased Aromatic–Aliphatic Copolyesters: Poly(Butylene Furandicarboxylate-Co-

ACS Paragon Plus Environment

30

Page 31 of 35 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

Succinate)s Obtained by Ring Opening Polymerization. Polym. Chem. 2017, 8 (4), 748– 760, DOI 10.1039/C6PY01879C. (33)

Ishii, N.; Inoue, Y.; Shimada, K.; Tezuka, Y.; Mitomo, H.; Kasuya, K. Fungal Degradation of Poly(Ethylene Succinate). Polym. Degrad. Stab. 2007, 92 (1), 44–52, DOI 10.1016/j.polymdegradstab.2006.09.014.

(34)

Kleeberg, I.; Hetz, C.; Kroppenstedt, R. M.; M??ller, R. J.; Deckwer, W. D. Biodegradation of Aliphatic-Aromatic Copolyesters by Thermomonospora Fusca and Other Thermophilic Compost Isolates. Appl. Environ. Microbiol. 1998, 64 (5), 1731– 1735.

(35)

Kleeberg, I.; Welzel, K.; VandenHeuvel, J.; Müller, R.-J.; Deckwer, W.-D. Characterization of a New Extracellular Hydrolase from Thermobifida Fusca Degrading Aliphatic−Aromatic Copolyesters. Biomacromolecules 2005, 6 (1), 262–270, DOI 10.1021/bm049582t.

(36)

Muroi, F.; Tachibana, Y.; Kobayashi, Y.; Sakurai, T.; Kasuya, K. Influences of Poly(Butylene Adipate-Co-Terephthalate) on Soil Microbiota and Plant Growth. Polym. Degrad. Stab. 2016, 129, 338–346, DOI 10.1016/j.polymdegradstab.2016.05.018.

(37)

Kasuya, K.; Ishii, N.; Inoue, Y.; Yazawa, K.; Tagaya, T.; Yotsumoto, T.; Kazahaya, J. ichiro; Nagai, D. Characterization of a Mesophilic Aliphatic-Aromatic CopolyesterDegrading Fungus. Polym. Degrad. Stab. 2009, 94 (8), 1190–1196, DOI 10.1016/j.polymdegradstab.2009.04.013.

ACS Paragon Plus Environment

31

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

(38)

Page 32 of 35

Suzuki, M.; Tachibana, Y.; Kazahaya, J.; Takizawa, R. Difference in Environmental Degradability between Poly ( Ethylene Succinate ) and Poly ( 3-Hydroxybutyrate ). J. Polym. Res. 2017, 24, 217, DOI 10.1007-s10965-017-1383-4.

(39)

Li, X.; Yee, A. F. Design of Mechanically Robust High-T g Polymers: Mechanical Properties of Glassy Poly(Ester Carbonate)s with Cyclohexylene Rings in the Backbone. Macromolecules 2004, 37 (19), 7231–7239, DOI 10.1021/ma0489568.

(40)

Yamadera, R.; Murano, M. The Determination of Randomness in Copolyesters by High Resolution Nuclear Magnetic Resonance. J. Polym. Sci. Part A-1 Polym. Chem. 1967, 5 (9), 2259–2268, DOI 10.1002/pol.1967.150050905.

(41)

Yamadera, R.; Murano, M. Studies on Tacticity of Polyacrylonitrile. I. High-Resolution Nuclear Magnetic Resonance Spectra of Polyacrylonitrile. J. Polym. Sci. Part A-1 Polym. Chem. 1967, 5 (5), 1059–1071, DOI 10.1002/pol.1967.150050511.

(42)

Arandia, I.; Mugica, A.; Zubitur, M.; Arbe, A.; Liu, G.; Wang, D.; Mincheva, R.; Dubois, P.; Müller, A. J. How Composition Determines the Properties of Isodimorphic Poly(Butylene Succinate- Ran -Butylene Azelate) Random Biobased Copolymers: From Single to Double Crystalline Random Copolymers. Macromolecules 2015, 48 (1), 43–57, DOI 10.1021/ma5023567.

(43)

Wood, L. A. Glass Transition Temperatures of Copolymers. J. Polym. Sci. 1958, 28 (117), 319–330, DOI 10.1002/pol.1958.1202811707.

(44)

Balachandar, M.; Balakrishnan, T.; Kothandaraman, H. Glass Transition and Melting Temperatures of Random Copolyesters of Poly(Ethylene Terephthalate) with p-

ACS Paragon Plus Environment

32

Page 33 of 35 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

Hydroxybenzoic Acid, 3. Die Makromol. Chemie 1983, 184 (2), 443–453, DOI 10.1002/macp.1983.021840218. (45)

Flory, P. J. Thermodynamics of Crystallization in High Polymers. IV. A Theory of Crystalline States and Fusion in Polymers, Copolymers, and Their Mixtures with Diluents. J. Chem. Phys. 1949, 17 (3), 223–240, DOI 10.1063/1.1747230.

(46)

Ichikawa, Y.; Kondo, H.; Igarashi, Y.; Noguchi, K.; Okuyama, K.; Washiyama, J. Crystal Structures of α and β Forms of Poly(Tetramethylene Succinate). Polymer (Guildf). 2000, 41 (12), 4719–4727, DOI 10.1016/S0032-3861(99)00659-X.

(47)

Pérez-Camargo, R. A.; Fernández-D’Arlas, B.; Cavallo, D.; Debuissy, T.; Pollet, E.; Avérous, L.; Müller, A. J. Tailoring the Structure, Morphology, and Crystallization of Isodimorphic Poly(Butylene Succinate-Ran-Butylene Adipate) Random Copolymers by Changing Composition and Thermal History. Macromolecules 2017, 50 (2), 597–608, DOI 10.1021/acs.macromol.6b02457.

(48)

Baba, T.; Tachibana, Y.; Suda, S.; Kasuya, K. Evaluation of Environmental Degradability Based on the Number of Methylene Units in Poly(Butylene n-Alkylenedionate). Polym. Degrad. Stab. 2017, 138, 18–26, DOI 10.1016/j.polymdegradstab.2017.02.007.

(49)

Shah, A. A.; Kato, S.; Shintani, N.; Kamini, N. R.; Nakajima-Kambe, T. Microbial Degradation of Aliphatic and Aliphatic-Aromatic Co-Polyesters. Appl. Microbiol. Biotechnol. 2014, 98 (8), 3437–3447, DOI 10.1007/s00253-014-5558-1.

(50)

Hisano, T.; Kasuya, K. I.; Tezuka, Y.; Ishii, N.; Kobayashi, T.; Shiraki, M.; Oroudjev, E.; Hansma, H.; Iwata, T.; Doi, Y.; et al. The Crystal Structure of Polyhydroxybutyrate

ACS Paragon Plus Environment

33

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 34 of 35

Depolymerase from Penicillium Funiculosum Provides Insights into the Recognition and Degradation of Biopolyesters. J. Mol. Biol. 2006, 356 (4), 993–1004, DOI 10.1016/j.jmb.2005.12.028. (51)

Gan, Z.; Kuwabara, K.; Abe, H.; Iwata, T.; Doi, Y. The Role of Polymorphic Crystal Structure and Morphology in Enzymatic Degradation of Melt-Crystallized Poly(Butylene Adipate) Films. Polym. Degrad. Stab. 2005, 87 (1), 191–199, DOI 10.1016/j.polymdegradstab.2004.08.007.

(52)

Díaz, A.; Franco, L.; Estrany, F.; del Valle, L. J. J.; Puiggalí, J. Poly(Butylene AzelateCo-Butylene Succinate) Copolymers: Crystalline Morphologies and Degradation. Polym. Degrad. Stab. 2014, 99 (1), 80–91, DOI 10.1016/j.polymdegradstab.2013.11.022.

ACS Paragon Plus Environment

34

Page 35 of 35 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

Synopsis A copolymerization with novel furfural-based building block improved the thermal and mechanical propertis of poly(butylene succinate).

Table of Contents

ACS Paragon Plus Environment

35