Feasibility of Solid-State Postpolymerization on Fossil- and Bio-Based

May 5, 2016 - Fossil-based and true bio-based poly(butylene succinate) (PBS) prepolymers were synthesized and submitted to solid-state polymerization ...
0 downloads 0 Views 911KB Size
Article pubs.acs.org/IECR

Feasibility of Solid-State Postpolymerization on Fossil- and BioBased Poly(butylene succinate) Including Polymer Upcycling Routes Constantine D. Papaspyrides,*,† Stamatina Vouyiouka,† Ioanna-Nektaria Georgousopoulou,† Sinisa Marinkovic,‡ Boris Estrine,‡ Catherine Joly,§ and Patrice Dole∥ †

Laboratory of Polymer Technology, School of Chemical Engineering, National Technical University of Athens, Zographou Campus, Athens 15780, Greece ‡ Green Chemistry Department, Agro-Industrie Recherches et Développements, Route de Bazancourt, Pomacle 51110, France § Laboratoire de Bioingénierie et Dynamique Microbienne aux Interfaces Alimentaires, IUT Lyon 1 site de Bourg en Bresse, Technopole Alimentec, Université de Lyon, Université Lyon 1-ISARA Lyon, rue Henri de Boissieu, Bourg en Bresse F-01000, France ∥ Centre Technique de la Conservation des Produits Agricoles, Technopole Alimentec, rue Henri de Boissieu, Bourg en Bresse F-01000, France ABSTRACT: Fossil-based and true bio-based poly(butylene succinate) (PBS) prepolymers were synthesized and submitted to solid-state polymerization (SSP) in the proximity of the polyester melting point (Tm), for reaction times up to 29 h under flowing nitrogen. SSP acted as a postcrystallization process, imparting an increase of the PBS melting point up to 126 °C from a starting Tm of 112−114 °C. Adding a precrystallization step prior SSP even resulted in a 2.5 times increase of the initial MW and a Tm shift up to 128 °C. Furthermore, the effect of most critical process parameters on the SSP feasibility and effectiveness was assessed, so as to launch an appropriate operation profile. End-group imbalance turned out to be the most significant key parameter for PBS polymerizability, and various attempts were made toward correcting it. Finally, SSP was examined as a PBS recycling technique and efficiently “revived” hydrolyzed PBS structures.



melt polycondensation process (Figure 1)13,16,24−28 with a first step of esterification between BDO and SA usually in a molar ratio of 1.1:1 at temperatures from 190 to 225 °C. After removal of the stoichiometric water and in the presence of an efficient catalyst, the formed oligomers are further polycondensed at higher temperatures of 230−250 °C under vacuum, resulting in high-molecular-weight PBS, for example, with a number-average molecular weight (M̅ n) of 120000 g mol−1.13,24−26 Typical catalysts include tin(II) chloride (SnCl2),27 tetrabutyl titanate (TBT),25,26,28 distannoxane,29−31 lanthanide triflates,32 p-toluenesulfonic acid,33 with the most common being TBT. However, melt polymerization products usually suffer from drawbacks associated with the high reaction temperatures and residence times in the melt, as well as yellowing caused by titanium-based catalysts.28,34−36 Especially in the case of PBS, high esterification temperatures for prolonged times and application of vacuum favor BDO loss (boiling point, Tb = 230 °C), its dehydration to tetrahydrofuran,10,13,16 and other side reactions, all resulting in a disturbed end-group balance

INTRODUCTION Nowadays, the negative environmental impact caused by the use and disposal of conventional nondegradable plastics has intensified the efforts of academia and industry toward the development of bio-based and biodegradable polymers.1−6 As a representative of these new materials, poly(butylene succinate) (PBS) is considered to be promising for the replacement of polyolefins, like polyethylene (PE) and polypropylene (PP).7 Compared to the most commercialized biopolymer [poly(lactic acid), PLA; Tg = 60 °C], the PBS lower glass transition point (Tg = −32 °C) renders it flexible, as well as more easily meltprocessable at lower cost.8−11 In addition, PBS possesses a relatively high melting point (Tm = 112−114 °C)12,13 compared to other soft polyesters derived from aliphatic diols and diacids,14 mechanical properties similar to those of PP and low-density PE,12 and a tunable structure via alteration of the monomer composition.8,15,16 Finally, the high potential for industrial production of partially or fully bio-based PBS has stimulated interest for its research and development; already ameliorated fermentation processes have given rise to the production of a bio-based diacid monomer (succinic acid, SA) at industrial scale,8,15−23 while the diol monomer, 1,4butanediol (BDO), can also be prepared from bio-based SA.8,15 A significant number of publications have dealt with the synthesis of PBS: it is conventionally prepared via a two-stage © 2016 American Chemical Society

Received: Revised: Accepted: Published: 5832

February 12, 2016 April 15, 2016 May 5, 2016 May 5, 2016 DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842

Article

Industrial & Engineering Chemistry Research

Figure 1. Common synthetic route of PBS via a two-step melt polycondensation process.

Table 1. Characteristics of the SSP Starting Materials Fossil-Based PBS F1 F2 F3 F4 F5 oligomer Bio-Based PBS B1

form

M̅ n (g mol−1)

M̅ w (g mol−1)

PDI

pellets flakes flakes flakes flakes flakes

34800 10400 16700 26200 46000 2022

68100 18000 28700 46500 85700 3892

1.96 1.74 1.72 1.78 1.86 1.92

61.42 143.45 357.26 159.35 101.36 58.64

pellets

33300

61800

1.86

111.86 ± 0.86

and impaired properties such as hydrolysis stability.37,38 These phenomena may even be more intense in bio-based PBS production; the organic acid-nature impurities, e.g., acetic acid, which are recovered from the fermentation broth together with SA,38,39 are likely to contribute to side reactions, lowering the obtainable maximum molecular weight (MW)37 and changing the polymer characteristics, such as thermal and mechanical ones.24 These problems can be circumvented by applying milder polymerization conditions. Already, in the work of Kong et al.,24 lower reaction temperatures (T < 200 °C) for 16 h in the absence of vacuum and chain extenders, were found to be beneficial to synthesizing PBS of M̅ n = 116000 g mol−1 and presenting less yellowing. Furthermore, enzymatic polymerization of PBS has also been examined using lipases as biocatalysts.40−43 An alternative technique under mild conditions is solid-state polymerization (SSP).44−54 SSP belongs to the bulk (solventfree) methods and involves MW-increasing reactions in the amorphous regions of the semicrystalline solid prepolymer where sufficient segmental and reactive species mobility exists. The reaction temperatures lie below the melting point of the starting material, while inert gas flow or vacuum is applied for the removal of polycondensation byproducts. SSP is industrially used to produce high-molecular-weight polyamides, such as PA 66, and other thermoplastic polyesters e.g., poly(ethylene terephthalate) (PET) and poly(butylene terephthalate).43−56 However, keeping the system in the solid state is not always achieved, and sintering may occur, although the operational temperature is set below the starting material melting point. Especially in the case of PAs, a distinct but strange transition from the solid-to-melt state (SMT) occurs, which is markedly favored at higher reaction temperatures. This SMT phenomenon has been extensively studied by Papaspyrides et al. and correlated to the accumulation of water formed during reactions, yielding in this way lower melting hydrated regions. 44−53 In order to avoid these implications, a precrystallization step is usually introduced prior to the main SSP process, as is widely used for PET57 and studied for PLA.58 Apart from the postpolymerization technique, SSP has also been industrially implemented as a route for recycling, and relevant systems exist, with important companies supporting the design and construction of plants.59,60 Especially for PET

[COOH] (mequiv kg−1) ± ± ± ± ± ±

3.15 1.23 0.32 0.61 0.38 2.93

Tm1 (°C) 113.2 115.5 115.2 116.6 117.4 103.1

± ± ± ± ± ±

0.8 0.7 0.2 0.0 0.2 0.5

115.7 ± 2.9

Td (°C) 393.7 ± 4.1 395.8 ± 4.0 395.7 ± 2.0 395.3 ± 0.5 399.1 ± 1.0 not measured 406.4 ± 3.1

bottle-to-bottle recycling and decontamination, SSP yields recycled products of improved quality in terms of MW, color, lower acetaldehyde content, and the absence of contaminants.61−66 Cruz et al.64 applied SSP on postconsumer PET and observed that the inherent viscosity of the SSP products reached the value of the virgin material (0.78 dL g−1) and reduced the carboxyl group concentration. Rieckmann et al.66 developed a model to describe a closed-loop recycling system for PET beverage bottles, which indicated that after a single recycling loop, all of the implicated quality parameters achieved the specifications via careful tuning of the temperature, residence time, and surface area for degassing. The application of SSP to bio-based and biodegradable polymers constitutes a challenging case study, with a limited number of relevant publications. SSP has been applied to PLA as a crystal reorganization and monomer-removal technique and as a finishing step,53,54 while recently a combination of enzymatic polymerization and SSP has been applied to PBS.14 SSP feasibility as a recycling approach has also been reported for PLA:56 hydrolyzed samples were subjected to SSP at temperatures of 2.5−25.0 °C below their melting point, and the process achieved an increase of up to 1.7 times the initial MW. However, despite these few works, there is still a literature gap on PBS SSP. This may be due to the lower operating temperatures ( [OH]I and [COOH]II < [OH]II}. The PBS oligomer was synthesized according to the procedure (Table 1), and the presence of OH-terminated species in the oligomer was verified in 1H NMR analysis, where a triplet at 3.65−3.69 ppm was attributed to −CH2OH (Figure 5). The oligomer was incorporated in grade B1 at a concentration of 7.2 wt % based on eq 6 (B1olig). The proper incorporation can be verified by the higher intensity of the −CH2OH triplet compared to the unmodified grade (B1), which was, however, lower than that of the oligomer. In addition, oligomer incorporation resulted in a lowering of the MW and an increase of PDI of B1, as anticipated (Table 5). On the other hand, the relevant DSC analysis (Figure 4c) showed that the oligomer endotherm peak was not apparent in grade B1olig; the latter implies that the new OH groups were mainly 5838

DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842

Article

Industrial & Engineering Chemistry Research Table 6. Characteristics of Hydrolyzed PBS Grades grade

M̅ n (g mol−1)

M̅ w (g mol−1)

[COOH] (mequiv kg−1)

IV (dL g−1)

Tm1 (°C)

ΔH1 (J g−1)

xc (%)

hF2 hF3 hB1

5500 4700 7640

11000 9400 15300

382.40 ± 3.31 718.04 ± 20.50 335.09 ± 8.76

0.18 ± 0.03 0.16 ± 0.03 0.24 ± 0.00

112.9 ± 0.6 112.4 ± 0.7 115.2 ± 0.5

86.0 ± 1.9 87.5 ± 1.6 79.8 ± 7.0

78 79 72

allocated in the amorphous regions and the crystalline phases remained almost intact. B1olig was then subjected to a two-step SSP (105 °C for 4 h + 24 h at 114 °C), revealing, however, that in this case the precrystallization step was not sufficient to prevent the sticking problems probably due to oligomer incorporation. On the basis of the SEC results (Table 5), no significant M̅ w buildup was evidenced, while M̅ n increased by 27% along with a decrease in PDI. This result might be explained by the type of polycondensation reactions that occurred in the presence of the oligomer. It could be said that the oligomer either reacted with PBS macromolecules without significantly extending them and/or participated in homocondensation reactions through transesterification. Homocondensation led to low-MW species crystallized in the amorphous regions of the prepolymer, as also evidenced by the DSC curve of B1olig (Figure 4b): two melting populations appeared after SSP, implying the existence of two different crystal qualities. Another possible scenario for the nonameliorated SSP behavior could be that the oligomers were not homogeneously distributed in the material because of their polarity or their markedly different MWs compared to the polymeric matrix. SSP of Hydrolyzed PBS (Recycling Approach). In order to examine SSP as a recycling technique, fossil-based grades F2 and F3 as well as bio-based grade B1 were submitted to accelerated hydrolysis runs. Table 6 summarizes the properties of the resultant hydrolyzed grades, with hydrolyzed fossil-based grade hF3 exhibiting a higher hydrolysis rate (ΔIV = −51%) compared to hydrolyzed fossil-based grade hF2 (ΔIV = −13%) because of the enhanced COOH content, which autocatalyzed hydrolysis reactions as anticipated.84 Hydrolyzed bio-based grade hB1 presented in total the highest drop of IV, 67%, probably because of the acid-nature impurities of bio-based SA, which exerted an autocatalysis effect and rendered the material more susceptible to hydrolysis. Turning to the DSC results, the effect of hydrolysis on the semicrystalline structure of PBS was noticeable. The melting points decreased (up to 3 °C), while exposure to water vapor plasticized the hydrolyzed PBS structure and rendered it more mobile and faster crystallizable, i.e., xc increased up to 79% from 69%. Scouting SSP runs were then carried out at various reaction times, using the resultant hydrolyzed samples, so as to assess whether the decreased MW and thermal properties can be repaired to some extent. Figure 6a illustrates the variation of IV as a function of the SSP reaction time, and it can be observed that the best results in terms of MW buildup were obtained after the completion of 24 h without the occurrence of sticking problems. Focusing on the two fossil-based grades, hF2 and hF3, the latter noted higher SSP effectiveness, reaching IV 0.32 ± 0.02 dL g−1 very close to the value of the starting material (grade F3). In parallel, [COOH] was respectively decreased by approximately 35% and 29% for hF3 and hF2. Bio-based grade hB1 also presented an increase of up to 27% IV and concurrently a carboxylic group content consumption of up to 22%. Concerning the thermal properties of hF2, hF3, and hB1, Figure 6b documents the upgrade of Tm through the SSP-based

Figure 6. Variation of the (a) intrinsic viscosity (b) melting point, and (c) degree of crystallinity versus SSP reaction time for hydrolyzed PBS grades. 5839

DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842

Article

Industrial & Engineering Chemistry Research

(3) Chen, H. B.; Wang, X. L.; Zeng, J. B.; Li, L. L.; Dong, F. X.; Wang, Y. Z. A Novel Multiblock Poly(ester urethane) Based on Poly(butylene succinate) and Poly(ethylene succinate-co-ethylene terephthalate). Ind. Eng. Chem. Res. 2011, 50, 2065. (4) Zeng, J. B.; Huang, C. L.; Jiao, L.; Lu, X.; Wang, Y. Z.; Wang, X. L. Synthesis and Properties of Biodegradable Poly(butylene succinateco-diethylene glycol succinate) Copolymers. Ind. Eng. Chem. Res. 2012, 51, 12258. (5) Kim, Y. Z.; Kang, G. D.; Yoon, K. C.; Park, O. O. Comparison of mechanical properties of blended and synthesized biodegradable polyesters. Macromol. Res. 2014, 22, 382. (6) Luo, X.; Li, J.; Feng, J.; Yang, T.; Lin, X. Mechanical and thermal performance of distillers grains filled poly(butylene succinate) composites. Mater. Eng. 2014, 57, 195. (7) Zhu, Q. Y.; He, Y. S.; Zeng, J. B.; Huang, Q.; Wang, Y. Z. Synthesis and characterization of a novel multiblock copolyester containing poly(ethylene succinate) and poly(butylene succinate). Mater. Chem. Phys. 2011, 130, 943. (8) Babu, R. P.; O'Connor, K.; Seeram, R. Current progress on biobased polymers and their future trends. Biomaterials 2013, 2, 8. (9) Papageorgiou, G. Z.; Bikiaris, D. N. Crystallization and melting behavior of three biodegradable poly(alkylene succinates). A comparative study. Polymer 2005, 46, 12081−12092. (10) Xu, J.; Guo, B. H. In Plastics from Bacteria: Natural Functions and Applications; Chen, G. Q., Ed.; Springer: Berlin, 2010; p 347; DOI: 10.1007/978-3-642-03287-5_14. (11) Kanemura, C.; Nakashima, S.; Hotta, A. Mechanical properties and chemical structures of biodegradable poly(butylene-succinate) for material reprocessing. Polym. Degrad. Stab. 2012, 97, 972. (12) Chrissafis, K.; Paraskevopoulos, K. M.; Bikiaris, D. N. Thermal degradation mechanism of poly(ethylene succinate) and poly(butylene succinate): Comparative study. Thermochim. Acta 2005, 435, 142. (13) Garin, M.; Tighzert, L.; Vroman, I.; Marinkovic, S.; Estrine, B. The kinetics of poly(butylene succinate) synthesis and the influence of molar mass on its thermal properties. J. Appl. Polym. Sci. 2014, 131, XX. (14) Kanelli, M.; Douka, A.; Vouyiouka, S.; Papaspyrides, C.; Topakas, E.; Papaspyridi, L. M.; Christakopoulos, P. Production of biodegradable polyesters via enzymatic polymerization and solid state finishing. J. Appl. Polym. Sci. 2014, 131 (19), XX. (15) Hwang, S. Y.; Yoo, E. S.; Im, S. S. The synthesis of copolymers, blends and composites based on poly(butylene succinate). Polym. J. 2012, 44, 1179. (16) Xu, J.; Guo, B. H. Poly(butylene succinate) and its copolymers: research, development and industrialization. Biotechnol. J. 2010, 5, 1149. (17) Yang, Q.; Hirata, M.; Hsu, Y. I.; Lu, D.; Kimura, Y. Improved thermal and mechanical properties of poly(butylene succinate) by polymer blending with a thermotropic liquid crystalline polyester. J. Appl. Polym. Sci. 2014, 131, XX. (18) Top value added chemicals from biomass Volume I-Results of screening for potential candidates from sugars and synthesis gas; U.S. Department of Energy: Washington, DC, 2004 (19) SUCCIPACK FP 7 Project, www.succipack.eu. (20) http://www.purac.com/en/sustainability/biobased-succinicacid.aspx. (21) http://www.bio-amber.com/products/en/products/succinic_ acid. (22) http://www.reverdia.com/wp-content/uploads/european_ technology_and_industry-news_sustainable_sucinnic_acid.pdf. (23) http://www.chemicals-technology.com/projects/myriant-plant/ . (24) Kong, X.; Qi, H.; Curtis, J. M. Synthesis and characterization of high-molecular weight aliphatic polyesters from monomers derived from renewable resources. J. Appl. Polym. Sci. 2014, 131, XX. (25) Bikiaris, D. N.; Achilias, D. S. Synthesis of poly(alkylene succinate) biodegradable polyesters I. Mathematical modelling of the esterification reaction. Polymer 2006, 47, 4851.

recycling process. In particular, already after 2 h, Tm was enhanced by almost 8 °C and, in parallel, xc noted a slight decrease, which varied between ca. 3% and 10% (Figure 6c). This behavior may be ascribed to the greater rejection of chain ends to the interlamellar regions, produced from hydrolytic degradation, and the short SSP time, which was not adequate for reorganization of the polymeric chains into thicker crystalline domains. Longer SSP times accordingly further improved the xc values, which reached up to ∼90%. Finally, regarding thermal degradation of the PBS SSP recyclates, in all cases the degradation temperature increased from 396 to 405 °C, returning to the value of the virgin material. Taking into account that above, it can be concluded that SSP comprises a promising repairing tool for hydrolyzed PBS structures and further work needs to be done so as to enhance even more the process effectiveness.



CONCLUSIONS The present work investigated for the first time the feasibility of SSP for the production of fossil-based and true bio-based PBS, working on two axes: SSP as an extension to PBS melt polymerization and as a recycling tool to “revive” at some extent the polyester hydrolyzed structure. SSP was found in both cases to effectively increase the MW and thermal properties of PBS: the most successful results were obtained for a two-step process for an operating temperature window fixed at approximately 3 °C below the prepolymer melting point. Apart from the reaction temperature, the most critical process parameter was found to be the end-group imbalance in the majority of prepolymers and the presence mainly of COOH-capped macromolecules. Attempts were made toward repairing the stoichiometry, with slightly however ameliorated results implying the need for milder melt polymerization conditions so as to avoid hydroxyl group consumption. Finally, the origin of PBS prepolymers (fossil- vs bio-based grades) revealed different SSP behavior and end product properties, such as sintering phenomena and multiple melting of bio-based SSP products. The latter results indicated the need to differentiate the SSP process when dealing with true biobased starting materials.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the EC within the 7th Framework Program under Grant 289196 (Development of active, intelligent and sustainable food Packaging using PolybutyleneSUCCInate-SUCCIPACK).



REFERENCES

(1) Nikolic, M.; Djonlagic, J. Synthesis and characterization of biodegradable poly(butylene succinate-co-butylene adipate)s. Polym. Degrad. Stab. 2001, 74, 263. (2) Zheng, L.; Li, C.; Huang, W.; Huang, X.; Zhang, D.; Guan, G.; Xiao, Y.; Wang, D. Synthesis of high-impact biodegradable multiblock copolymers comprising of poly(butylene succinate) and poly(1,2propylene succinate) with hexamethylene diisocyanate as chain extender. Polym. Adv. Technol. 2011, 22, 279. 5840

DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842

Article

Industrial & Engineering Chemistry Research

(49) Papaspyrides, C. D. Solid state polyamidation. In The polymeric materials encyclopedia; Salamone, J. C., Ed.; CRC Press, Inc.: Boca Raton, FL, 1996; pp 7819−7831. (50) Vouyiouka, S.; Karakatsani, E.; Papaspyrides, C. Solid state polymerization. Prog. Polym. Sci. 2005, 30 (1), 10. (51) Papaspyrides, C. D.; Vouyiouka, S. N.; Bletsos, I. V. New aspects on the mechanism of the solid state polyamidation of PA 6,6 salt. Polymer 2006, 47, 1020. (52) Papaspyrides, C., Vouyiouka, S., Eds. Solid state polymerization; John Wiley & Sons, Inc.: Hoboken, NJ 2009. (53) Vouyiouka, S.; Papaspyrides, C. Solid state polymerization. Encyclopedia of Polymer Science and Technology, 4th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, 2011; pp1−32. (54) Zhang, J.; Shen, X. J.; Zhang, J.; Feng, L. F.; Wang, J. J. Experimental and modeling study of the solid state polymerization of poly(ethylene terephthalate) over a wide range of temperatures and particle sizes. J. Appl. Polym. Sci. 2013, 127, 3814. (55) Kim, Y. J.; Kim, J.; Oh, S. G. Modeling of solid-state polymerization of bisphenol A polycarbonate. Ind. Eng. Chem. Res. 2012, 51, 2904. (56) Vouyiouka, S.; Theodoulou, P.; Symeonidou, A.; Papaspyrides, C. D.; Pfaendner, R. Solid state polymerization of poly(lactic acid): some fundamental parameters. Polym. Degrad. Stab. 2013, 98, 2473. (57) Kim, T. Y.; Lofgren, E. A.; Jabarin, S. A. Solid-state polymerization of poly(ethylene terephthalate). I. Experimental study of the reaction kinetics and properties. J. Appl. Polym. Sci. 2003, 89, 197. (58) Peng, B.; Hou, H.; Song, F.; Wu, L. Synthesis of high molecular weight poly(l-lactic acid) via melt/solid state polycondensation. II. Effect of precrystallization on solid state polycondensation. Ind. Eng. Chem. Res. 2012, 51, 5190. (59) www.buhlergroup.com. (60) Wadekar, S.; Agarwal, U.; Boon, W.; Nadkarni, V. In Solid state polymerization; Papaspyrides, C, Vouyiouka, S, Eds.; John Wiley & Sons: Hoboken, NJ, 2009; Chapter 8. (61) Thiele, U. 4th China International Recycled Polyester Fiber Market & Tech Forum, Hangzou, People’s Republic of China, Sept 2− 4, 2008. (62) Welle, F. Twenty years of PET bottle to bottle recycling − An overview. Res. Conserv. Rec. 2011, 55, 865. (63) Awaja, F.; Pavel, D. Recycling of PET. Eur. Polym. J. 2005, 41, 1453. (64) Cruz, S. A.; Zanin, M. PET recycling: Evaluation of the solid state polymerization process. J. Appl. Polym. Sci. 2006, 99, 2117. (65) Welle, F. Simulation of the Decontamination Efficiency of PET recycling processes based on solid-state polycondensation. Packag. Technol. Sci. 2013, 27, 141. (66) Rieckmann, T.; Frei, F.; Volker, S. Modelling of PET quality parameters for a closed-loop recycling system for food contact. Macromol. Symp. 2011, 302, 34. (67) Solomon, O. F.; Ciuta, I. Z. Détermination de la viscosité intrinsèque de solutions de polymères par une simple détermination de la viscosité. J. Appl. Polym. Sci. 1962, 6, 683. (68) Bogdanic, G. Group contribution methods for estimating the properties of polymer systems. Hem. Ind. 2006, 60, 287. (69) Culbert, B.; Christel, A. Continuous solid state polycondensation of polyesters. In Modern polyesters: Chemistry and technology of polyesters and copolyesters; Scheirs, J., Long, T., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, 2003. (70) Vouyiouka, S. N.; Filgueiras, V.; Papaspyrides, C. D.; Lima, E. L.; Pinto, J. C. Morphological changes of poly(ethylene terephthalateco-isophthalate) during solid state polymerization. J. Appl. Polym. Sci. 2012, 124, 4457. (71) Yasuniwa, M.; Tsubakihara, S.; Satou, T. J.; Iura, K. Multiple melting behavior of poly(butylene succinate). II. Thermal analysis of isothermal crystallization and melting process. J. Polym. Sci., Part B: Polym. Phys. 2005, 43, 2039.

(26) Bikiaris, D. N.; Achilias, D. S. Synthesis of poly(alkylene succinate) biodegradable polyesters, Part II: Mathematical modelling of the polycondensation reaction. Polymer 2008, 49, 3677. (27) Zhu, C.; Zhang, Z.; Liu, Q.; Wang, Z.; Jin, J. Synthesis and biodegradation of aliphatic polyesters from dicarboxylic acids and diols. J. Appl. Polym. Sci. 2003, 90, 982. (28) Jacquel, N.; Freyermouth, F.; Fenouillot, F.; Rousseau, A.; Pascault, J. P.; Fuertes, P.; Saint-Loup, R. Synthesis and properties of poly(butylene succinate): Efficiency of different transesterification catalysts. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 5301. (29) Velmathi, S.; Nagahata, R.; Sugiyama, J. I.; Takeuchi, K. A. A rapid eco-friendly synthesis of poly(butylene succinate) by a direct polyesterification under microwave irradiation. Macromol. Rapid Commun. 2005, 26, 1163. (30) Takahashi, H.; Hayakawa, T.; Ueda, M. Convenient synthesis of poly(butylene succinate) catalyzed by distannoxane. Chem. Lett. 2000, 29, 684. (31) Ishii, M.; Okazaki, M.; Shibasaki, Y.; Ueda, M.; Teranishi, T. Convenient Synthesis of aliphatic polyesters by distannoxane-catalyzed polycondensation. Biomacromolecules 2001, 2, 1267. (32) Takasu, A.; Oishi, Y.; Iio, Y.; Inai, Y.; Hirabayashi, T. Synthesis of aliphatic polyesters by direct polyesterification of dicarboxylic acids with diols under mild conditions catalyzed by reusable rare-earth triflate. Macromolecules 2003, 36, 1772. (33) Song, D. K.; Sung, Y. K. Synthesis and characterization of biodegradable poly(1,4-butanediol succinate). J. Appl. Polym. Sci. 1995, 56, 1381. (34) Medellin-Rodriguez, F. J.; Lopez-Guillen, R.; Waldo-Mendoza, M. A. Solid-state polymerization and bulk crystallization behavior of poly(ethylene terephthalate) (PET). J. Appl. Polym. Sci. 2000, 75, 78. (35) Filgueiras, V.; Vouyiouka, S. N.; Papaspyrides, C. D.; Lima, E. L.; Pinto, J. C. Solid-state polymerization of poly(ethylene terephthalate): the effect of water vapor in the carrier gas. Macromol. Mater. Eng. 2011, 296, 113. (36) Duh, B. Solid-state polymerization of poly(trimethylene terephthalate). J. Appl. Polym. Sci. 2003, 89, 3188. (37) Ye, Y.; Choi, K. Y. Optimizing polymer reactivities for the solidstate polycondensation of AA and BB type monomers. Polymer 2008, 49, 2817. (38) de Gooijer, J. M.; Scheltus, M.; Jansen, M. A. G.; Koning, C. E. Carboxylic acid end group modification of poly(butylene terephalate) in supercritical fluids. Polymer 2003, 44, 2201. (39) Kurzrock, T.; Weuster-Botz, D. Recovery of succinic acid from fermentation broth. Biotechnol. Lett. 2010, 32, 331. (40) An, S.; Zhu, J.; Lu, D.; Liu, Z.; Xuebao, H. Lipase-catalyzed synthesis and characterization of high-molecular-weight PBS. CIESC Journal 2013, 64, 1855. (41) Linko, Y. Y.; Wang, Z. L.; Seppala, J. Lipase-catalyzed synthesis of poly(1,4-butyl sebacate) from sebacic acid or its derivatives with 1,4butanediol. J. Biotechnol. 1995, 40, 133. (42) Liu, W.; Chen, B.; Wang, F.; Tan, T.; Deng, L. Lipase-catalyzed synthesis of aliphatic polyesters and properties characterization. Process Biochem. 2011, 46, 1993. (43) McKinlay, J. B.; Vieille, C.; Zeikus, J. G. Prospects for a biobased succinate industry. Appl. Microbiol. Biotechnol. 2007, 76, 727. (44) Papaspyrides, C.; Kampouris, E. Solid-state polyamidation of dodecamethylenediammonium adipate. Polymer 1984, 25, 791. (45) Papaspyrides, C.; Kampouris, E. Influence of metal catalysts on solid state polyamidation of nylon salts. Polymer 1986, 27, 1437. (46) Papaspyrides, C. Solid state polyamidation of nylon salts. Polymer 1988, 29, 114. (47) Papaspyrides, C. Solid state polyamidation of aliphatic diaminealiphatic diacid salts: a generalized mechanism for the effect of polycondensation water on reaction behavior. Polymer 1990, 31 (3), 490. (48) Papaspyrides, C. Solid state polyamidation processes. Polym. Int. 1992, 29, 293. 5841

DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842

Article

Industrial & Engineering Chemistry Research (72) Yasuniwa, M.; Satou, T. J. Multiple melting behavior of poly(butylene succinate). I. Thermal analysis of melt-crystallized samples. J. Polym. Sci., Part B: Polym. Phys. 2002, 40, 2411. (73) Georgousopoulou, I. N.; Vouyiouka, S.; Dole, P.; Papaspyrides, C. D. Thermo-mechanical degradation and stabilization of poly(butylene succinate). Polym. Degrad. Stab. 2016, 128, 182. (74) Wang, C. D.; Guo, B.; Li, R. Synthesis, characterization, and properties of long-chain branched poly(butylene succinate). J. Appl. Polym. Sci. 2012, 124, 1271. (75) Katiyar, V.; Shaama, M. S.; Nanavati, H. J. A comprehensive single-particle model for solid-state polymerization of poly(L-lactic acid). J. Appl. Polym. Sci. 2011, 122 (5), 2966. (76) Bashir, Z.; Al-Aloush, I.; Al-Raqibah, I.; Ibrahim, M. J. Evaluation of three methods for the measurement of crystallinity of PET resins, preforms and bottles. Polym. Eng. Sci. 2000, 40, 2442. (77) Young, R. J.; Lovell, P. A. Introduction to polymers; CRC Press, Taylor & Francis Group: Boca Raton, FL, 2011. (78) Hiemenz, P. C.; Lodge, T. P. Polymer Chemistry; CRC Press, Taylor & Francis Group: Boca Raton, FL, 2007. (79) Hu, L.; Wu, L.; Song, F.; Li, B. G. Kinetics and modeling of melt polycondensation for synthesis of poly[(butylene succinate)-co(butylene terephthalate)], 1-Esterification. Macromol. React. Eng. 2010, 4, 621. (80) McKinlay, J. B.; Vieille, C.; Zeikus, J. G. Prospects for a biobased succinate industry. Appl. Microbiol. Biotechnol. 2007, 76, 727. (81) Song, H.; Lee, S. Y. Production of succinic acid by bacterial fermentation. Enzyme Microb. Technol. 2006, 39, 352. (82) Zeikus, J. G.; Jain, M. K.; Elankovan, P. Biotechnology of succinic acid production and markets for derived industrial products. Appl. Microbiol. Biotechnol. 1999, 51, 545. (83) Varadarajan, S.; Miller, D. J. Catalytic upgrading of fermentation-derived organic acids. Biotechnol. Prog. 1999, 15, 845. (84) Antheunis, H.; van der Meer, J. C.; de Geus, M.; Heise, A.; Koning, C. O. Autocatalytic equation describing the change in molecular weight during hydrolytic degradation of aliphatic polyesters. Biomacromolecules 2010, 11, 1118.

5842

DOI: 10.1021/acs.iecr.6b00588 Ind. Eng. Chem. Res. 2016, 55, 5832−5842