Renewable Unsaturated Polyesters from Muconic Acid - ACS

Sep 27, 2016 - cis,cis-Muconic acid is an unsaturated dicarboxylic acid that can be produced in high yields via biological conversion of sugars and li...
1 downloads 11 Views 1MB Size
Subscriber access provided by TUFTS UNIV

Article

Renewable unsaturated polyesters from muconic acid Nicholas Andrew Rorrer, John Robert Dorgan, Derek R. Vardon, Chelsea R Martinez, Yuan Yang, and Gregg T Beckham ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.6b01820 • Publication Date (Web): 27 Sep 2016 Downloaded from http://pubs.acs.org on September 30, 2016

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 free 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 accessible to all readers and 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.

ACS Sustainable Chemistry & Engineering 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 31

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

Renewable unsaturated polyesters from muconic acid Nicholas A. Rorrer,a,b John R. Dorgan,b Derek R. Vardon,a Chelsea R. Martinez,a Yuan Yang,c Gregg T. Beckhama,* a. National Bioenergy Center, 15013 Denver West Drive, National Renewable Energy Laboratory, Golden CO 80401 b. Department of Chemical and Biological Engineering, Colorado School of Mines, 1500 Illinois Street, Golden CO 80401 c. Chemistry Department, 1500 Illinois Street, Colorado School of Mines, Golden, CO 80401 Biochemicals | lignocellulose | polymers | lignin valorization Corresponding Author * [email protected] KEYWORDS: Unsaturated Polyester, Composites, Muconic Acid, Poly(butylene succinate)

ABSTRACT: cis,cis-Muconic acid is an unsaturated dicarboxylic acid that can be produced in high yields via biological conversion of sugars and lignin-derived aromatic compounds. Muconic acid is often targeted as an intermediate to direct replacement monomers such as adipic or terephthalic acid. However, the alkene groups in muconic acid provide incentive for its direct use in polymers, for example, in the synthesis of unsaturated polyester resins. Here, biologically derived muconic acid is incorporated into polyesters via condensation polymerization using the homologous series of poly(ethylene succinate), poly(propylene succinate), poly(butylene succinate), and poly(hexylene succinate). Additionally, dimethyl 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 2 of 31

cis,cis muconate is synthesized and subsequently incorporated into poly(butylene succinate). NMR measurements demonstrate that alkene bonds are present in the polymer backbones. In all cases, the glass transition temperatures are increased while the melting and degradation temperatures are decreased. In the case of poly(butylene succinate), utilization of neat muconic acid yields sub-stoichiometric incorporation consistent with a tapered copolymer structure, while the muconate diester exhibits stoichiometric incorporation and a random copolymer structure based on thermal and mechanical properties. Prototypical fiberglass panels were produced by infusing a mixture of low molecular weight poly(butylene succinate-co-muconate) and styrene into a woven glass mat and thermally initiating polymerization resulting in thermoset composites with shear moduli in excess of 30 GPa, a value typical of commercial composites. The increased glass transition temperatures with increasing mucconic incorporation leads to improved composites properties. We find that the molecular tunability of poly(butylene succinate-co-muconate) as a tapered or random copolymer enables the tunability of composite properties. Overall, this study demonstrates the utility of muconic acid as a monomer suitable for direct use in commercial composites.

Introduction Lignocellulose is a promising renewable feedstock for producing a wide variety of monomers and polymers.1–5 In an integrated biorefinery, similar to a petroleum refinery, coproduction of monomers for polymer materials along with biofuels can improve profitability.6 Through various routes under intense development, biomass can be fractionated into carbohydrates and lignin and transformed into fuel, chemicals, and material building blocks.7–10 To do so, upgrading processes that incorporate direct biological, chemo-catalytic, or hybrid technologies are needed.11–13 Importantly, the high oxygen content of biomass (~35-45%) also makes it ideal for targeting oxygen-rich platform molecules such as polyester precursors with high atom efficiency when compared to petroleum to which oxygen must be added.14

ACS Paragon Plus Environment

Page 3 of 31

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

cis,cis-Muconic acid is a promising dicarboxylic acid renewable platform molecule that can be accessed from both sugars15 and lignin16 and be readily converted to both terephthalic acid and adipic acid, which can further be used to produce poly(ethylene terephthate) (PET) or Nylon 6,6, respectively. In the case of terephthalic acid, trans,trans-muconic acid is isomerized to cis,cis muconic acid and subsequently reacted with ethylene and dehydrated, while adipic acid can be produced directly from the hydrogenation of muconic acid.16–19 While the use of muconic acid as a precursor to direct replacement monomers is being thoroughly explored, exploitation of its alkene bonds for utilization as a functional replacement in polymers has received less attention. As the use of muconic acid without further chemical processing has potential for economic and environmental process advantages, there is incentive to explore utilization of muconic acid as a monomer for polymer applications. One potential application of muconic acid is unsaturated polyester (UPE) resins, which are versatile materials used in various applications with a growing global market forecast to reach USD 10.48 billion in 2019.20,21 UPEs form durable structures and coatings when cross-linked with a reactive monomer, such as styrene. Molecularly tunable UPEs are highly desirable; for example, the ability to adjust thermal properties such as the glass transition, melting, and degradation temperatures is considered advantageous.

UPEs are ideal candidates for incorporating bioderived materials. Specifically, their synthesis can be readily achieved using unsaturated dicarboxylic acids and dialcohols, both of which are readily produced via biochemical routes.22–25 While there has been considerable technical effort aimed at replacing styrene in these widely used materials,26 there are few reports of producing the unsaturated polyester from renewable resources.27,28 Furthermore, reports to date utilize itaconic acid as the source of renewable unsaturation, thereby producing a polyester chain in which the unsaturated double bonds are peripheral to the chain backbone. Interestingly, a recent study utilized a partially hydrogenated muconic acid to provide a single unsaturated bond into a polyamide.29

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 4 of 31

This works details the use of bio-derived cis,cis muconic acid and dimethyl cis,cis muconate in combination with other bio-derivable monomers to synthesize linear biobased polymers for UPEs. Demonstrating the synthesis of these new materials based on cis,cis muconic acid opens the pathway for its use as a functional replacement of high-value petroleum-derived materials (Figure 1), which will improve both the sustainability metrics of the materials and improve biorefinery economics. In the present study, a family of copolymers is produced by incorporating muconic acid into four different succinatebased polyesters. The properties of these copolymers differ from their corresponding homopolymers. In addition, a series of low molecular weight copolymer poly(butylene succinate-co-muconate) (PBSM) are synthesized (from neat cis,cis muconic acid and dimethyl cis,cis muconate), combined with styrene, infused into a long glass fiber mat, and crosslinked to create a prototypical fiberglass panel. Overall, the present study conclusively demonstrates the value of bio-derived muconic acid in producing novel UPEs. Lignocellulosic Biomass

Bio-Derived Monomer O OH HO O

cis,cis-Muconic Acid

O

O

O OH HO

HO

HO

OH

O

R

OH

OH HO

Succinic Acid

-Diol

H

O

O R1

O O

R2

HO

O

O

R2

OH

R

Resin

Functional Replacements

H N

N H

O

O H

Composite

H

O

Nylon 6,6

O

Unsaturated Polyester

TPA

O

O

O

O

O

Adipic Acid

O

cis,cis-Muconic Acid

OH

HO

OH

O O

PET

Direct Replacements

Figure 1. Select examples of the conversion of muconic acid to direct or functional replacements. Muconic acid produced from sugars or lignin-derived aromatic compounds can be catalytically converted to existing commercial monomer units, and subsequently implemented in traditional materials. However,

ACS Paragon Plus Environment

Page 5 of 31

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

due to the additional functionality in muconic acid, new materials can be manufactured to serve as functional replacements to existing technologies. Experimental (Materials and Methods) Materials: Succinic acid, butanediol, ethylene glycol, propanediol, hexanediol, chloroform, methanol, azobisisobutyronitrile and titanium (IV) butoxide were purchased from Sigma Aldrich. To ensure high purity, succinic acid was recrystallized from acetic acid. All dialcohols were purified by vacuum distillation. Purity was confirmed by measuring melting points using DSC to be greater than 99.8% for all monomers.

Biological production of muconic acid: cis,cis-muconic acid was produced biologically with Pseudomonas putida, KT2440-CJ102,16,19 using benzoate as the precursor substrate. Muconic acid was produced at a titer of 34.5 g/L, and the separations process resulted in recovery of 28.1 g/L. Purity, as determined via DSC, was found to be greater than 99.5%. Further details can be found in our previous work.19

Polymerization: Polymerizations were conducted in three-necked round bottom flasks with an overhead Arrow 750 stirring motor. The three-necked round bottom flask was connected to a Dean-Stark trap connected to a water-cooled condenser and vacuum line. Initially the reactor was charged with the reactants at a molar ratio of 1:1.1 succinic acid:dialcohol. In the case of copolymer synthesis, the reactants were charged into the reaction at a ratio of x:(1-x):1 muconic acid or ester:succinic acid:dialcohol, where x is a fraction between 0 and 0.25. The temperature of the reaction vessel was raised to 150°C for two hours under a N2 purge. Oligomers were formed during the first two hours of reaction. At two hours, titanium (IV) butoxide catalyst was added to the reaction mixture at 0.1 wt% and vacuum was applied to bring the total pressure below 50 torr. The reactor temperature was then raised to 220°C and the reaction was allowed to proceed for an additional ten hours. The catalyst promotes the condensation reaction of the 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 6 of 31

monomers and transesterification between chains, thus favoring development of the most probable distribution. Investigations of additional catalysts for polymer transesterification were beyond the scope of this study, and will be examined in future work. After ten hours, the motor was stopped and the reaction was allowed to cool to room temperature under vacuum. The polymer was dissolved in chloroform and subsequently precipitated using a twofold volume of methanol. The precipitate was filtered and dried in a vacuum oven for 48 hours.

Dimethyl cis,cis-muconate synthesis: Dimethyl cis,cis-muconate was synthesized via the base-catalyzed method proposed by Frost et al.30 To do so, muconic acid was suspended in aqueous sodium hydroxide with dimethyl sulfate and stirred constantly for six (6) hours. Subsequently, the dimethyl cis,cis muconate was extracted with ethyl acetate, and then reextracted with sodium hydroxide and sodium chloride. Yields were comparable to the published work (52 wt.%).30

Fiberglass Composite: The composite material was produced from the 25% loaded poly(butylene succinate-co-muconate) with styrene as the crosslinking component. Azobisisobutyronitrile (AIBN) was used as a photoinitiator and woven fiberglass (Bondo brand) was purchased. The resin mixture comprised of 50 wt. % unsaturated polyester, 48.5 wt% styrene, 1 wt% Sartomer 350, and 0.5 wt% AIBN. Standard practice was followed when selecting the catalyst for this study; other initiators will be investigated in future work. The resin mixture was purged with N2 for five minutes and subsequently applied to the two pieces of fiberglass. The resin-soaked fiberglass sheets were placed between two Teflon sheets on a hot plate heated to 80°C and allowed to react for 4 hours. Subsequently the composite was removed from the mold and placed in a vacuum oven for 24 hours to remove trace unreacted styrene.

Thermal Measurements: Thermal properties were ascertained using a TA Instruments Q-5000 Digital Scanning Calorimeter (DSC) with aluminum hematic pans and using a TA Instruments Q-500 Thermal ACS Paragon Plus Environment

Page 7 of 31

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

Gravimetric Analyzer (TGA) with platinum pans. All DSC and TGA scans were run at a heating rate of 10°C/min unless otherwise noted.

Molecular Weight Determination: Molecular weights and polydispersities were ascertained using a Wyatt Instruments gel permeation chromatography system with light scattering detection for absolute molecular weight measurement. A constant value of the specific refractive index increment of (dn/dc) = 0.05 mL/g was adopted. HPLC grade chloroform was used as the carrier solvent.

NMR: Structure was determined via NMR using a JEOL ECA 500 MHz liquid-state NMR spectrometer. All

13

C NMR spectra were recorded under proton decoupling with 2s relaxation delay. For both

homopolymers and copolymers, deuterated chloroform was mixed with deuterated dimethyl sulfoxide and used in a volume ratio of 2:1 for the solvent.

FTIR: IR Spectra was obtained via diffuse-reflectance Fourier-transform infrared spectroscopy using a Themo Nicolet iS50 FT-IR spectrometer operating at 4 cm-1 resolution at room temperature under nitrogen equipped with a Harrick Praying Mantis controlled-environment chamber and KBr windows.

Mechanical Testing: Storage and Loss moduli for the polymers were determined using the single cantilever fixture on a TA Instruments Q5000 Dynamic Mechanical Analyzer (DMA). Rectangular samples having dimensions of 35 x 5.0 x 1.0 mm were prepared using a Hakke MiniJet Extruder. Modulus values for the composite material were determined using a Rheometric ARES-LS Rheometer using torsional rectangular geometry fixtures. A rectangular sample having dimensions of 12.5 x 5.0 x 2.0 mm was cut from the sheet using an Isomet diamond saw.

Results 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 31

To investigate the effect of muconic acid incorporation on polymer properties, a series of homopolymers was first synthesized and characterized. Poly(butylene succinate) (PBS), poly(ethylene succinate) (PES), poly(propylene succinate) (PPS), and poly(hexylene succinate) (PHS) were synthesized via catalyzed condensation polymerization. Scheme 1 illustrates the condensation with transesterification polymerization reaction in the presence of titanium butoxide IV as a catalyst to form the copolymers; in the case of the homopolymer, there is no muconic acid present. Homopolymer structures were confirmed using one-dimensional 1H and 13C NMR spectroscopy. The molecular weight distribution was determined by GPC. GPC traces of the homopolymers are provided in Supporting Information Figure S1. A representative NMR spectrum for PBS is provided in Figure 2A. O

O excess

+

OH

HO

HO

OH

+

OH

HO O

O Condensation Oligomerization O O

O

HO

O

OH

O

O O

O m

n Transesterification O O H O

O O

OH O

O

O n

m

Scheme 1. The condensation-transesterification reaction implemented in this work for polyester synthesis. Initially, an excess of the dialcohol is reacted with the diacid mixture to form oligomers (in this case succinic and cis,cis muconic acid). Subsequently, a transesterfication catalyst is introduced to form high molecular weight polymers.

ACS Paragon Plus Environment

Page 9 of 31

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

Chloroform O

bc

HO

d

O

O

DMSO

A

H

a

O

a d

b

c

O O

f O HO

O

O

O

c

e

O

b

a

a

B

OH

O

d

d

b

Chloroform

DMSO

c f 190

180

170

160

xx 180 xx 160

Figure 2.

13

e 150

140

xx 140

130

120

110

100

90

80

xx 120 xx 100 xx 80

70

xx

60

50

60 xx

40

40

30

20

xx 20

ppm

ppm

C NMR of the PBS homopolymer (A) and the muconic acid-containing copolymer

PBSM(A)-12.8% (B) Incorporation of the double bond is evidenced by signals between 120 to 160 ppm. Homopolymer Synthesis: Thermal properties for all synthesized homopolymers were determined using differential scanning calorimetry (DSC) and thermal gravimetric analysis (TGA) under N2. The glass transition temperature, Tg, the melting temperature, Tm, and the TD,50, the temperature at which 50% mass loss occurs, were measured for each homopolymer. Both thermal and molecular weight data for the homopolymers are summarized in Table 1. PBS is the most robust homopolymer based on its favorable thermal properties (Tg ~ -40°C, Tm = 120°C, TD,50 = 425°C). In contrast to PBS, the other polymers exhibit melting temperatures lower than 100°C, usually around 60°C. However, all polymers exhibit glass transition temperatures below -10°C and degradation temperatures above 400°C. The higher molecular weight of the poly(butylene succinate) is indicative of the higher reactivity of succinic acid over muconic acid. All of the polymers exhibit similar enthalpies of fusion.

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 10 of 31

Table 1. Homopolymer thermal and molecular weight properties. Thermal properties determined by DSC and TGA and weight average molecular weight determined via GPC with light scattering detection. Tm (°C)

Tg (°C)

TD,50 (°C)

∆Hf (J/g)

Mw

Ð

Poly(ethylene succinate)

105

-13

445

36.7

4.7*104

1.9+/-0.1

Poly(propylene succinate)

48

-34

415

43.1

6.1*104

1.9+/-0.1

Poly(butylene succinate)

130

-31

425

64.1

1.95*105

2.0+/-0.1

Poly(hexylene succinate)

58

-40

430

52.7

8.7*104

2.0 +/- 0.1

Muconic Acid Incorporation: Neat cis,cis muconic acid was subsequently incorporated into each of the polymers to form UPEs. To do so, cis, cis-muconic acid was added to the reaction mixture at an initial loading of 10 mol% of the total carboxylic groups. NMR spectra confirming the incorporation, albeit at a reduced level compared to that charged to the reactor, are shown in the Supporting Information (Figure S3). Additionally, a representative FTIR spectra is presented in the Supporting Information (Figure S4). However, due to the low incorporation of muconic acid and the strong C=O and O-H stretches NMR is implemented to provided conclusive evidence for muconic incorporation. As an illustrative example and given the superior thermal properties and emerging commercial significance of PBS,31 muconic acid was also added to PBS reaction mixtures at 1, 5 and 25 mol% of the initial acid groups to form a series of copolymers that exhibit varying amounts of unsaturation. The polymers formed are poly(butylene succinate-co-muconate), abbreviated as PBSM(A/E)-XX%, where XX denotes the incorporation of muconic acid into the polymer and the A/E designation represents if the polymer was synthesized with cis,cis muconic acid (A) or dimethyl cis,cis muconate (E). The NMR spectra again confirm the presence of double bonds in the copolymer backbone. A representative spectrum for the 25% muconic acid loaded material is presented in Figure 2B; signals “e” and “f” correspond to the two distinct alkene carbons, which are not present in the PBS homopolymer. After synthesis, all polymers were dissolved and reprecipitated to remove unreacted monomers. Accordingly, the NMR spectra can be considered conclusive

ACS Paragon Plus Environment

Page 11 of 31

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

evidence of muconic acid incorporation. In all cases using cis,cis muconate, quantitative analysis by peak integration demonstrates that the final level of incorporation is lower than the initial muconic acid charge to the reactor. Finally, the NMR spectra indicate that the muconic structural unit maintains the cis,cis configuration; there is no evidence of thermally induced isomerization; if isomerization had taken place, additional peaks would be present. H-NMR is presented in Figure S2.

Based on the results above, it can be inferred that muconic acid is less reactive than the other saturated diacids present in the reaction mixtures. That is, the NMR results show incorporation below the stoichiometry of the reaction mixture. This lack of reactivity may be attributable to steric effects but it is likely that the delocalized electrons provided by the double bonds decreases the acidity relative to a saturated structure. The reduced reactivity is also evident in the GPC results where total molecular weight decreases with increasing muconic acid charge in the reactor. GPC chromatograms are presented in Figure S3.

Muconic acid incorporation leads to copolymers with thermal properties distinct from the corresponding homopolymers. As an example of these changes, Figure 3 shows the DSC trace for the second heating of PBS and PBSM(A)-12.8%. These data show an emergence of a second Tg, a phenomenon present at all levels of muconic acid incorporation in PBS that differ in magnitude Tg and location of emergence depending on the extent of muconic acid incorporation. This effect is further demonstrated by the trend of decreasing Tm with increasing muconic acid content shown in Figure 3. This melting point depression phenomena is reminiscent of eutectic-like behavior, which has been observed in other copolyesters.32–34 The lower glass transition temperatures increase as the number of muconic acid groups in the copolymer increases. This increase can be rationalized along the lines of well-known structure-property relationships: namely, the alkene bonds inhibit bond rotation in the polymer backbone, and are expected to retard segmental motion, thus increasing Tg. At the same time, the structural dissimilarity of the muconic group ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

disrupts crystalline packing, leading to less perfect crystals and therefore decreased Tm. The emergence of two glass transition temperatures in PBS likely indicates a tapered copolymer structure.35–37 5

∆T = 50 °C/min

A

B

Second Tg Region

3

Heat Flow, W/g

4

Heat Flow, W/g

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 31

2

3 10

20

30

40

50

60

70

2

1 1

PBSM(A)-12.8% PBSM(A)-6.4% PBSM(A)-2.2%

PBS PBSM(A)-12.8% 0

0 -50

0

50

100

150

0

25

50

75

100

125

150

175

Temperature, °C

Temperature, °C

Figure 3. (A) Differential scanning calorimetry scan at an elevated heating rate of 50°C/min for poly(butylene succinate) homopolymer and poly(butylene succinate-co-muconate) copolymer at 25% loading. The elevated ramp rate is used to elucidate the two glass transition temperatures present in the copolymer, as indicated by the dotted black lines. (B) Differential scanning calorimetry scans for poly(butylene succinate-co-muconate) copolymers at a scan rate of 50°C/min. The inset shows the melting temperature depression and the shift in the glass transition temperature that occurs with increasing incorporation.

TGA results for the PBSM copolymers are presented in Figure 4. As expected, incorporation of reactive double bonds leads to decreased thermal stability. Increasing the amount of unsaturation in the polymer backbone leads to lower TD,50 values. Both thermal and molecular weight data for the copolymers are summarized in Table 2.

ACS Paragon Plus Environment

Page 13 of 31

1.0

0.8

Mass Fraction, φ

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

0.6

0.4

0.2

PBS PBSM(A)-2.2% PBSM(A)-6.4% PBSM(A)-12.8%

0.0 200

400

600

Temperature, °C

Figure 4. TGA traces of poly(butylene succinate-co-muconate) melts demonstrating decreasing degradation temperature with increasing incorporation.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 14 of 31

Table 2. Copolymer muconic incorporation, thermal properties, and weight average molecular weight with polydispersity index (Ð). Initially muconic acid is loaded as 10 mol% of the total acid, or 5 mol% of total monomer concentration. Muconic acid incorporation is reported as a function of total polymer concentration.

Homopolymer

Muconic Muconic Acid Tm Tg Tg,2 TD,50 ∆Hf Acid Loading Incorporation (°C) (°C) (°C) (°C) (J/g) (mol% acid) (mol % acid)

Poly(ethylene succinate)

10

Poly(propylene succinate)

10

6.4

Poly(butylene succinate)

1

--

Poly(hexylene succinate)

4.6

Mw

Ð

--

420

66.2

4.0*104

1.9 +/- 0.1

103

-9

46

-31

--

410

47.3

5.7*104

1.9 +/- 0.1

-30

--

425

65.2 1.92*105

2.0 +/- 0.1

129 5

2.2

125

-29

45

410

69.7 1.21*105

2.0 +/- 0.1

10

6.4

123

-18

42

390

73.2 1.03*105

1.9 +/- 0.1

25

12.8

105

-13

39

380

74.7

9.7*104

1.9 +/- 0.1

10

5.6

46

-31

--

415

58.4

7.3*104

1.9 +/- 0.1

Esterification of Muconic Acid for Polymerization: Esterification of the diacid monomer is not typically required for catalyzed polymerization reactions with saturated monomers. However, diester monomers are commonly used in the commercial production of PET to circumvent the poor solubility and high melting temperature of terephthalic acid. To explore the utility of using the diester, cis,cis muconic acid was esterified under basic conditions to produce dimethyl cis,cis muconate. The muconate ester is more soluble in the diol (and subsequently the reaction mixture) than the neat muconic acid and has a lower melting point; these attributes are summarized in Table 3. Initially, dimethyl cis,cis muconate comprising 12.5% of the total non-diol concentration was charged to the ACS Paragon Plus Environment

Page 15 of 31

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

reactor and subsequently polymerized in the previously described manner. Importantly, utilization of the diester yields a polymer with near stoichiometric incorporation of muconate and a slightly higher molecular weight. A comparison of the monomer incorporation is provided in the Supporting Information (Table S1). Table 3. Properties of cis,cis muconic acid, dimethyl cis,cis muconate, and the copolymer synthesized with dimethyl cis,cis muconate. Property

Value

cis, cis Muconic Acid Melting Point

197°C

Purity

99.2 mol.%

Solubility in Butanediol (mol:mol)

0.05:1

Dimethyl cis,cis Muconate Melting Point

78°C

Purity

99.1%

Solubility in Butanediol

1:1

(mol:mol)

PBSM(E)-12.5% Ester Incorporation Muconate Incorporation (mol % acid units)

12.4 %

Molecular Weight

1.86*105 g/mol

Ð

2.0 +/- 0.1

Melting Temperature, TM

115 °C

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering

Glass Transition, Tg

-15 °C

Decomposition Temperature, TD,50

385°C

Figure 5 provides thermal data for the PBSM(E)-12.5% synthesized with the dimethyl ester as the starting material. As before, the melting temperature is depressed, the glass transition temperature is elevated, and the decomposition temperature is lowered. This is consistent with previous experiments: muconic acid is incorporated into the backbone of the polymer, resulting in retarded segmental motion and disruption of crystalline packing. However, unlike previous experiments with the PBSM copolymers, there is no second Tg. This lack of the second Tg combined with anincrease in the Tg and stoichiometric incorporation of the muconate indicates that the structure of PBSM synthesized with the ester is that of a random copolymer.35–37 5 1.0

PBSM-12.5% (Ester)

4

3

Mass Fraction, φ

0.8

Heat Flow, W/g

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 31

0.6

0.4

0.2

0.0 200

300

400

500

600

Temeprature, °C

2

1

0 -50

0

50

100

150

Temeprature, °C

Figure 5. Thermal data for the PBSM(E)-12.5% polymer synthesized using dimethyl cis,cis muconate ester. In the DSC scan, there is no evidence of a second glass transition.

ACS Paragon Plus Environment

16

Page 17 of 31

Mechanical Properties of Poly(Butylene Succinate-co-muconate) from cis,cis Muconic Acid and dimethyl cis,cis Muconate: To evaluate polymer performance and to confirm the presence of a second Tg in the PBSM(A) copolymers, dynamic mechanical analysis (DMA) was performed on injection-molded pieces of the copolymer. Figure 6 provides the storage and loss modulus for the PBSM copolymers. Overall, the copolymers exhibit a higher modulus than that of PBS (E’ ~ 300 MPa at 35°C) and demonstrate behavior that is expected of polyesters above their first glass transition temperature.38 The second Tg in the samples synthesized with neat cis,cis muconic acid is indicated by an inflection in the storage modulus followed by a local maximum in the loss modulus. The location of the moduli is higher than the Tg reported via DSC, which is consistent with theory and

1000

100

750

75

500

50

250

25

PBSM(A)-12.8% PBSM(A)-6.4% PBSM(A)-2.2% PBSM(E)-12.4%

0

Loss Modulus, MPa

other experimental observations.39

Storage Modulus, MPa

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 40

50

60

70

80

90

Temperature, °C

Figure 6. Storage and loss modulus for the PBSM copolymer as a function of temperature. The inflection in the storage modulus and local maximum in the loss modulus is indicative of a glass transition temperature. The second glass transition temperature decreases with increasing muconic incorporation, in agreement with the observation from DSC. For comparison, the PBSM(E) does not show an inflection or local maximum.

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering

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

Page 18 of 31

In the case of the copolymer synthesized from the ester (PBSM(E)-12.4%), there is no inflection in the storage modulus, nor is there a local maximum in the loss modulus, indicating that there is no second Tg, indicating that the copolymer is random when synthesized with the ester. The copolymer still possesses a higher modulus than PBS, however, the PBSM(E)-12.4% possesses a slightly lower modulus than the polymer synthesized with neat acid. This is consistent with the aforementioned hypothesis that when PBSM is synthesized with the acid it possesses a tapered structure, while in the case of the ester the structure, it forms a random copolymer.

Fiber Glass Composite: To demonstrate the applicability of these polymers, a series of composites were made with PBSM-12.5% (both acid- and ester-synthesized versions), styrene, a diacrylate crosslinking agent (Sartomer 350), and AIBN (azobisisobutyronitrile) as an initiator. This typical UPE resin mixture was applied to double ply fiberglass and pressed between two Teflon sheets at 80°C. The formulation is given in Table 4 and the reaction mechanism for cross-linking is depicted in Scheme 2. The mechanical properties of the composite were measured by DMA over a range of frequencies. Figure 7 shows that the modulus is independent of frequency for all of the composites, demonstrating that the composites behave as elastic solids rather than a viscoelastic polymer over the frequency range investigated. In the studied cases, the composites that were synthesized with the acid had higher moduli than the composites synthesized with the ester. Furthermore, in the case of the acid-synthesized material, the modulus of the final composite is found to decrease with increasing molecular weight of the precursor UPE. This indicates that the PBSM(A) copolymers provide tunability due to their unique structure. The highest measured shear modulus in this study was 31.8 GPa, which is typical of fiberglass composites formed using UPEs. This composite exhibited a glass transition temperature of 90°C, a degradation temperature of 400°C, and no Tm.

ACS Paragon Plus Environment

18

Page 19 of 31

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

O O

O

O

60 wt%

+ 40 wt%

O O O

R

O O

O Crosslinking with AIBN

R

O

O

O O

O O

O O

R R

O

O O

O O

O O

R

O

Scheme 2. Reaction scheme for crosslinking poly(succinate-co-muconate) in which a monomer with a terminal vinyl group is reacted with the polymer to form a cross-linked polymer network. In this work, the crosslinking unit is styrene.

Table 4. Biobased UPE fiberglass composite properties Property

Value

Resin Formulation PBSM(A)-12.8%

57.2 parts per hundred resin (pph)

Styrene

41.0 pph

Diacrylate

0.8 pph

AIBN

1.0 pph

Composite Properties Fiberglass Loading

68 wt.%

Resin Loading

32 wt.%

PBSM(A)-12.8% Molecular Weight

3.7*103 g/mol

Shear Modulus

31.8 GPa

Glass Transition, Tg

90°C

Decomposition Temperature, TD,50

400°C

ACS Paragon Plus Environment

19

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

Shear Modulus, Pa

ACS Sustainable Chemistry & Engineering

10

Page 20 of 31

10

PBSM(A)-12.8%, Mn = 3.7*103 PBSM(A)-12.8%, Mn = 4.6*10

3

PBSM(E)-12.4%, Mn = 3.4*103 10-2

10-1

100

101

Frequency, ω (Hz)

Figure 7. Shear modulus as a function of frequency for the synthesized composite. Photo of the composite (lower right) on top of a sheet of woven fiberglass. Standard deviations are calculated from multiple composites synthesized at different times with the same formulation. Fiber content within a data set and across all data sets is found to be ±1.3 wt. %

Discussion This work focuses on the synthesis of bio-derivable UPEs and their application in composites. In the composite formulation, the UPE accounts for 60% of the total polymer weight, while the resin (in most cases styrene) comprises the remainder of the polymer weight. Recent work has investigated replacing styrene with methacrylates.26,40 Additionally, other work has demonstrated the possibility of producing styrene via biological conversion of glucose.41,42 Therefore through the combination of the muconic copolymers and either bio-derived styrene or an alternative, a completely bio-derived resin can be realized, which we will report in future studies.

ACS Paragon Plus Environment

20

Page 21 of 31

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

cis,cis-Muconic acid can be produced via the oxidative 1,2-cleavage of catechol, a central intermediate in degradation of aromatic compounds in many aromatic-catabolizing microbes.43 Moreover, muconic acid can be accessed from both sugars15 and lignin-derived aromatic compounds.16,44 Currently, muconic acid is not industrially produced at scale, but it is a promising platform chemical, primarily motivated by its facile conversion to adipic acid.16,19,45

Because muconic acid possesses terminal carboxylic acids, it is a strong candidate for direct polymerization and implementation in UPEs. It should be noted that muconic acid has poor solubility in all of the dialcohols used in this study. Furthermore it does not melt within the range of normal reaction conditions (Tm of 195°C). During preliminary experiments, a reaction mixture of pure muconic acid and any diol underwent degradation. Despite these limitations, muconic acid can be copolymerized at lower loadings. The initial homopolymer suite for muconic incorporation was selected because the structural units can be derived from molecules that can also be produced utilizing biochemical pathways. As such, it is anticipated that the family of polymers represented in this study could potentially be fully derived from renewable resources.

PBS was chosen as the most promising polymer for the inclusion of varying levels of muconic acid. PBS is commercially available from bioderived substrates (e.g., starch-based dextrose). Succinic acid is produced biologically and transformed to butanediol through the catalytic conversion of succinic acid.46,47 Additionally, butanediol can be produced directly at high yields and titers.48 Furthermore, PBS has the most robust thermal profile of the homopolymers studied — namely, it exhibits a low Tg and a high Tm. This combination is useful in materials requiring ductility, toughness, and some thermal resistance.

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering

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

Page 22 of 31

Scheme 3. (A) Activation of succinic acid carbonyl carbon by Ti(OBu)4 catalyst (B) Corresponding activation of muconic acid, where charge is delocalized by the pi electron system. In all the polymers synthesized with neat cis,cis muconic acid, denoted PBSM(A), the resultant incorporation of muconic acid is lower than the initial stoichiometry, and the PBSM(A) copolymers possess a lower molecular weight and modified thermal degradation profiles. The sub-stoichiometric incorporation and the likely tapered molecular architecture can be attributed to the low reactivity of the muconic acid monomer. Namely, we hypothesize that the conjugated nature of muconate results in a greater distribution of electron density over the whole molecule, making it relatively less electrophilic at the carbonyl carbon compared with succinic acid (Scheme 3). This implies the carbon is less susceptible to nucleophilic attack by butanediol when complexed with the catalyst. In addition to the lowered reactivity, muconic acid is observed to have limited solubility in the reaction mixture (see Table 3).

Generally, the reaction conditions and assumption of equal reactivity of equivalent functional groups implies that all copolymers generated should be random copolymers. In fact, in the cases of PESM, PPSM, and PHSM there is no evidence of any non-random chain structure — a single Tg is

ACS Paragon Plus Environment

22

Page 23 of 31

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

observed. In the case of the PBSM(A) copolymers, two Tg are observed and both exhibit compositional dependence. The possibility of two separate molecular structures being present is discounted based on the GPC results (presented in Table 2 and SI Figure S1) and the TGA results (presented in Figure 4). Both indicate that only one molecular species is present. A separate elution peak or a broader molecular weight distribution (Ð > 2) in the GPC trace would indicate separate polymer structures. In addition, when multiple species are present, a TGA typically shows step-wise degradation regions. While speculative, the presence of two glass transitions might be attributed to the muconic acid groups reacting, on average, at a later stage of the polymerization reaction than the succinic acid; this would lead to the muconic acid-derived structural units being preferentially located nearer to the chain ends. Based on the relative, hypothesized reactivities of succinic and muconic acids, it is likely that the structure of the polymer would be block-like, which would give rise to a second Tg.

The dimethyl ester offers several advantages to neat muconic acid. Specifically, the insensitivity of transesterification to the presence of double bonds combined with increased solubility in the dialcohol and the lower melting point of the ester enables stoichiometric incorporation of muconate into the final polymer. This full incorporation enables more easily targeted polymer compositions. However, because the yield on the esterification reaction is 52%, overall utilization of the muconic acid substrate is equal. In addition, an increased molecular weight is obtained with the same reaction time if the muconate ester is compared to the neat muconic acid. The dimethyl cis,cis muconate possess the lowest melting temperature (all other forms of muconic acid and muconate possessing Tm in excess of 170 °C) making it the ideal candidate for polymerization. While the ester does offer the benefit of stoichiometric incorporation, there is an apparent benefit in the tapered structure of the

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 31

acid-derived polymers. Specifically, the tapered copolymer structure leads to a higher modulus in both the injection-molded polymers and composites. In the case of the PBSM(A) copolymer, there is a dependence of the modulus on molecular weight. As the molecular weight is increased, the modulus decreases, which implies an increase in the molecular weight between crosslinks for the composite. This effect could arise because the muconic acid units likely exist near the chain ends of the PBSM(A) copolymer. This control of the polymer molecular architecture enables tailoring of composite properties. In the case of the PBSM(E), however, the muconic units are likely randomly distributed through the polymer. Additionally, these changes in composite properties can be attributed to the unreacted muconic blocks stiffening the composite matrix. Due to the uniformity in crosslink site density, there is no inherent tunability present with the PBSM(E) copolymer and a lower modulus is obtained at the same muconic acid loading. It is important to note that the changes in modulus do not arise due to the change in fiber content (± 1.3 wt%), nor from the reactions yield (between 87% to 93%), as both variations are found to be non-correlative. Overall, the composites exhibit both a robust mechanical and thermal profile. The shear modulus of all synthesized composites is suitable for typical fiberglass composite, while the high glass transition temperature of 90°C and lack of melting temperature enables continual use up to the decomposition temperature. This proof-of-concept testing conclusively demonstrates that muconic acid can be utilized in the development of sustainable, tuneable, bio-renewable UPEs.

Conclusion As demonstrated in this work, muconic acid is a value-added, renewable monomer useful for producing bio-based crosslinkable resins. The presence of double bonds in the monomer also adds rigidity to the polymer backbone, thereby affecting its thermal properties. Several variants of the

ACS Paragon Plus Environment

24

Page 25 of 31

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

PBSM copolymer were used to fabricate industrially relevant UPEs having a shear modulus typical of fiberglass composites. Additionally, in the case of PBS incorporation, the molecular architecture of the polymer, and subsequent properties of the composite, can be tuned by selecting neat cis,cis muconic acid or dimethyl cis,cis muconate during synthesis. The substitution of fully renewable polymers in the ubiquitous use of unsaturated polyesters holds promise for significant environmental and economic benefits.

ASSOCIATED CONTENT Supporting Information. A document providing 13C NMR Spectra for the PBSM(A) melts, 1H NMR Spectra for PBSM(A)-12.8%, and GPC traces is provided in the supplemental information. The following files are available free of charge. AUTHOR INFORMATION Funding Sources Support for NAR was partially provided by the U.S. Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists, Office of Science Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education for the DOE under contract number DE-AC05-06OR23100. NAR, DRV, and GTB thank the NREL Laboratory Directed Research and Development program for partially funding the polymerization work. We thank the US Department of Energy Bioenergy Technologies Office for funding the muconate production efforts. CRM thanks the Colorado College Riley Scholar-inResidence program for research support. ACKNOWLEDGMENT

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

Page 26 of 31

We thank Davinia Salvachúa for producing the muconate used in this work and Christopher Johnson for use of the P. putida CJ102 strain. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes. ABBREVIATIONS PBS – Poly(butylene succinate), PES – Poly(ethylene succinate), PPS – Poly(propylene succinate), PHS – Poly(hexylene succinate). PBSM(A/E)-X.X% - Poly(butylene succinate-co-muconate) synthesized from the Acid (A) or Ester (E), NMR – Nuclear Magnetic Resonance, GPC – Gel Permeation Chromatography, DSC – Digital Scanning Calorimetry, TGA – Thermogravometric Analysis, DMA – Dynamic Mechanical Analysis, UPE – Unsaturated Polyester, Ð - PDI

ACS Paragon Plus Environment

26

Page 27 of 31

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

REFERENCES (1)

Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41 (4), 1538–1558.

(2)

Bozell, J. J.; Petersen, G. R. Technology development for the production of biobased products from biorefinery carbohydrates—the US Department of Energy’s “Top 10” revisited. Green Chem. 2010, 12 (4), 539.

(3)

Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107 (6), 2411–2502.

(4)

Werpy, T.; Petersen, G. Top value added chemicals from biomass. Volume I - Results of screening for potential candidates from sugars and synthesis gas; US Dep Energy, 2004.

(5)

Holladay, J.; Bozell, J.; White, J.; Johnson, D. Top value-added chemicals from biomass Volume II - Results of screening for potential candidates from biorefinery lignin; US Dep Energy, 2007.

(6)

Biddy, M.; Davis, R.; Tao, L.; Dowe, N.; Guarnieri, M. T.; Linger, J. G.; Karp, E. M.; Salvachua, D.; Vardon, D. R.; Beckham, G. T. The techno-economic basis for coproduct manufacturing to enable hydrocarbon fuel production from lignocellulosic biomass. ACS Sustain. Chem. Eng. 2016, 4 (6), 3196–3211.

(7)

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.

(8)

Zakzeski, J.; Bruijnincx, P. C.; Jongerius, A. L.; Weckhuysen, B. M. The catalytic valorization of lignin for the production of renewable chemicals. Chem. Rev. 2010, 110 (6), 3552–3599.

(9)

Ragauskas, A. J.; Beckham, G. T.; Biddy, M. J.; Chandra, R.; Chen, F.; Davis, M. F.; Davison, B. H.; Dixon, R. A.; Gilna, P.; Keller, M.; et al. Lignin valorization: Improving lignin processing in the biorefinery. Science 2014, 344 (6185), 1246843.

(10) Chundawat, S. P. S.; Beckham, G. T.; Himmel, M. E.; Dale, B. E. Deconstruction of lignocellulosic biomass to fuels and chemicals. Annu. Rev. Chem. Biomol. Eng. 2011, 2 (1), 121–145. (11) Xiong, M.; Deng, J.; Woodruff, A. P.; Zhu, M.; Zhou, J.; Park, S. W.; Li, H.; Fu, Y.; Zhang, K. A bio-catalytic approach to aliphatic ketones. Sci. Rep. 2012, 2. (12) Linger, J. G.; Vardon, D. R.; Guarnieri, M. T.; Karp, E. M.; Hunsinger, G. B.; Franden, M. A.; Johnson, C. W.; Chupka, G.; Strathmann, T. J.; Pienkos, P. T.; et al. Lignin valorization through integrated biological funneling and chemical catalysis. Proc. Natl. Acad. Sci. 2014, 111 (33), 12013–12018. (13) Anbarasan, P.; Baer, Z. C.; Sreekumar, S.; Gross, E.; Binder, J. B.; Blanch, H. W.; Clark, D. S.; Toste, F. D. Integration of chemical catalysis with extractive fermentation to produce fuels. Nature 2012, 491 (7423), 235–239. (14) Dusselier, M.; Mascal, M.; Sels, B. F. Top Chemical Opportunities from Carbohydrate Biomass: A Chemist’s View of the Biorefinery; Topics in Current Chemistry; Springer Berlin Heidelberg, 2014; pp 1–40.

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 31

(15) Draths, K. M.; Frost, J. W. Environmentally compatible synthesis of adipic acid from Dglucose. J. Am. Chem. Soc. 1994, 116 (1), 399–400. (16) Vardon, D. R.; Franden, M. A.; Johnson, C. W.; Karp, E. M.; Guarnieri, M. T.; Linger, J. G.; Salm, M. J.; Strathmann, T. J.; Beckham, G. T. Adipic acid production from lignin. Energy Environ. Sci. 2015, 8, 617–628. (17) Frost, J. W.; Draths, K. M. Synthesis of adipic acid from biomass-derived carbon sources. US5487987 A, January 30, 1996. (18) Lu, R.; Lu, F.; Chen, J.; Yu, W.; Huang, Q.; Zhang, J.; Xu, J. Production of diethyl terephthalate from biomass-derived muconic acid. Angew. Chem. Int. Ed. 2016, 55 (1), 249– 253. (19) Vardon, D. R.; Rorrer, N. A.; Salvachua, D.; Settle, A. E.; Johnson, C. W.; Menart, M. J.; Cleveland, N. S.; Ciesielski, P. N.; Steirer, K. X.; Dorgan, J.; et al. cis,cis-Muconic acid: Separation and catalysis to bio-adipic acid for nylon-6,6 polymerization. Green Chem. 2016, 18, 3397–3413. (20) Unsaturated Polyester Resin Market - Global Industry Analysis 2013 - 2019 by Transparency Market Research. (21) Reportlinker. Unsaturated Polyester Resin - Asia Pacific Industry Analysis, Size, Share, Growth, Trends and Forecast 2015 - 2023 http://www.prnewswire.com/newsreleases/unsaturated-polyester-resin---asia-pacific-industry-analysis-size-share-growth-trendsand-forecast-2015---2023-300211023.html (accessed Feb 10, 2016). (22) Harwood, C. S.; Parales, R. E. The β-ketoadipate pathway and the biology of self-identity. Annu. Rev. Microbiol. 1996, 50 (1), 553–590. (23) Chen, G.-Q. A microbial polyhydroxyalkanoates (PHA) based bio- and materials industry. Chem. Soc. Rev. 2009, 38 (8), 2434–2446. (24) Bechthold, I.; Bretz, K.; Kabasci, S.; Kopitzky, R.; Springer, A. Succinic acid: A new platform chemical for biobased polymers from renewable resources. Chem. Eng. Technol. 2008, 31 (5), 647–654. (25) Zeikus, J. G.; Jain, M. K.; Elankovan, P. Biotechnology of succinic acid production and markets for derived industrial products. Appl. Microbiol. Biotechnol. 1999, 51 (5), 545–552. (26) Cousinet, S.; Ghadban, A.; Fleury, E.; Lortie, F.; Pascault, J.-P.; Portinha, D. Toward replacement of styrene by bio-based methacrylates in unsaturated polyester resins. Eur. Polym. J. 2015, 67, 539–550. (27) Mohanty, A. K.; Misra, M.; Hinrichsen, G. Biofibres, biodegradable polymers and biocomposites: An overview. Macromol. Mater. Eng. 2000, 276-277 (1), 1–24. (28) Upton, B. M.; Kasko, A. M. Strategies for the conversion of lignin to high-value polymeric materials: Review and perspective. Chem. Rev. 2015, 116 (4), 2275–2306. (29) Suastegui, M.; Matthiesen, J. E.; Carraher, J. M.; Hernandez, N.; Rodriguez Quiroz, N.; Okerlund, A.; Cochran, E. W.; Shao, Z.; Tessonnier, J.-P. Combining metabolic engineering and electrocatalysis: Application to the production of polyamides from sugar. Angew. Chem. Int. Ed. 2016, 55 (7), 2368–2373.

ACS Paragon Plus Environment

28

Page 29 of 31

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

(30) Frost, J. W.; Miermont, A.; Schweitzer, D.; Bui, V. Preparation of trans, trans muconic acid and trans, trans muconates. US8426639 B2, April 23, 2013. (31) Xu, J.; Guo, B.-H. Poly(butylene succinate) and its copolymers: Research, development and industrialization. Biotechnol. J. 2010, 5 (11), 1149–1163. (32) Chen, C.-H.; Lu, H.-Y.; Chen, M.; Peng, J.-S.; Tsai, C.-J.; Yang, C.-S. Synthesis and characterization of poly(ethylene succinate) and its copolyesters containing minor amounts of butylene succinate. J. Appl. Polym. Sci. 2009, 111 (3), 1433–1439. (33) Nikolic, M. S.; Djonlagic, J. Synthesis and characterization of biodegradable poly(butylene succinate-co-butylene adipate)s. Polym. Degrad. Stab. 2001, 74 (2), 263–270. (34) Ahn, B. D.; Kim, S. H.; Kim, Y. H.; Yang, J. S. Synthesis and characterization of the biodegradable copolymers from succinic acid and adipic acid with 1,4-butanediol. J. Appl. Polym. Sci. 2001, 82 (11), 2808–2826. (35) Kim, J.; Mok, M. M.; Sandoval, R. W.; Woo, D. J.; Torkelson, J. M. Uniquely Broad Glass Transition Temperatures of Gradient Copolymers Relative to Random and Block Copolymers Containing Repulsive Comonomers. Macromolecules 2006, 39 (18), 6152–6160. (36) Gray, M. K.; Zhou, H.; Nguyen, S. T.; Torkelson, J. M. Differences in enthalpy recovery of gradient and random copolymers of similar overall composition: styrene/4-methylstyrene copolymers made by nitroxide-mediated controlled radical polymerization. Polymer 2004, 45 (14), 4777–4786. (37) Daimon, H.; Okitsu, H.; Kumanotani, J. Glass Transition Behaviors of Random and Block Copolymers and Polymer Blends of Styrene and Cyclododecyl Acrylate. I. Glass Transition Temperatures. Polym. J. 1975, 7 (4), 460–466. (38) Shibata, M.; Inoue, Y.; Miyoshi, M. Mechanical properties, morphology, and crystallization behavior of blends of poly(L-lactide) with poly(butylene succinate-co-L-lactate) and poly(butylene succinate). Polymer 2006, 47 (10), 3557–3564. (39) Wunderlich, B. The nature of the glass transition and its determination by thermal analysis. In Assignment of the Glass Transition; Seyler, R., Ed.; ASTM International: 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959, 1994; pp 17–17 – 15. (40) He, S. L.; Yaszemski, M. J.; Yasko, A. W.; Engel, P. S.; Mikos, A. G. Injectable biodegradable polymer composites based on poly(propylene fumarate) crosslinked with poly(ethylene glycol)-dimethacrylate. Biomaterials 2000, 21 (23), 2389–2394. (41) Claypool, J. T.; Raman, D. R.; Jarboe, L. R.; Nielsen, D. R. Technoeconomic evaluation of bio-based styrene production by engineered Escherichia coli. J. Ind. Microbiol. Biotechnol. 2014, 41 (8), 1211–1216. (42) Lian, J.; McKenna, R.; Rover, M. R.; Nielsen, D. R.; Wen, Z.; Jarboe, L. R. Production of biorenewable styrene: utilization of biomass-derived sugars and insights into toxicity. J. Ind. Microbiol. Biotechnol. 2016, 43 (5), 595–604. (43) Fuchs, G.; Boll, M.; Heider, J. Microbial degradation of aromatic compounds - from one strategy to four. Nat. Rev. Microbiol. 2011, 9 (11), 803–816.

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 31

(44) Beckham, G. T.; Johnson, C. W.; Karp, E. M.; Salvachúa, D.; Vardon, D. R. Opportunities and challenges in biological lignin valorization. Curr. Opin. Biotechnol. 2016, 42, 40–53. (45) Van de Vyver, S.; Román-Leshkov, Y. Emerging catalytic processes for the production of adipic acid. Catal Sci Technol 2013, 3 (6), 1465–1479. (46) Hong, U. G.; Park, H. W.; Lee, J.; Hwang, S.; Kwak, J.; Yi, J.; Song, I. K. Hydrogenation of succinic acid to 1,4-butanediol over rhenium catalyst supported on copper-containing mesoporous carbon. J. Nanosci. Nanotechnol. 2013, 13 (11), 7448–7453. (47) Kang, K. H.; Hong, U. G.; Bang, Y.; Choi, J. H.; Kim, J. K.; Lee, J. K.; Han, S. J.; Song, I. K. Hydrogenation of succinic acid to 1,4-butanediol over Re–Ru bimetallic catalysts supported on mesoporous carbon. Appl. Catal. Gen. 2015, 490, 153–162. (48) Yim, H.; Haselbeck, R.; Niu, W.; Pujol-Baxley, C.; Burgard, A.; Boldt, J.; Khandurina, J.; Trawick, J. D.; Osterhout, R. E.; Stephen, R.; et al. Metabolic engineering of Escherichia coli for direct production of 1,4-butanediol. Nat. Chem. Biol. 2011, 7 (7), 445–452.

ACS Paragon Plus Environment

30

Page 31 of 31

FOR TABLE OF CONTENTS USE ONLY

Acid Synthesized

Shear Modulus, Pa

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

Ester Synthesized

1010

O OH HO O

Muconate Incorporated Unsaturated Polyester Resins 10-2

10-1

100

101

Frequency, ω (Hz)

Renewable unsaturated polyesters from muconic acid Nicholas A. Rorrer, John R. Dorgan, Derek R. Vardon, Chelsea R. Martinez, Yuan Yang, Gregg T. Beckham

SYNOPSIS: Bio-catalytically derived cis, cis muconic acid is utilized as a functional replacement in unsaturated polyester resins.

ACS Paragon Plus Environment

31