You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/255763013

Advanced chemical recycling of poly(ethylene terephthalate) through


organocatalytic aminolysis

Article  in  Polymer Chemistry · February 2013


DOI: 10.1039/C2PY20793A

CITATIONS READS

80 792

11 authors, including:

Kazuki Fukushima Di S. Wei


The University of Tokyo University of Denver
67 PUBLICATIONS   3,480 CITATIONS    1 PUBLICATION   80 CITATIONS   

SEE PROFILE SEE PROFILE

Hans W Horn Gavin O Jones


IBM IBM
60 PUBLICATIONS   19,293 CITATIONS    55 PUBLICATIONS   3,000 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Natural fiber reinforcements and natural fillers for polymers View project

Quantum Chemistry with Quantum Computers View project

All content following this page was uploaded by Gavin O Jones on 05 November 2014.

The user has requested enhancement of the downloaded file.


Polymer
Chemistry
View Article Online
PAPER View Journal | View Issue

Advanced chemical recycling of poly(ethylene


terephthalate) through organocatalytic aminolysis†
Cite this: Polym. Chem., 2013, 4, 1610

Kazuki Fukushima,*a Julien M. Lecuyer,ab Di S. Wei,c Hans W. Horn,a Gavin O. Jones,a


Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

Hamid A. Al-Megren,d Abdullah M. Alabdulrahman,d Fares D. Alsewailem,d


Melanie A. McNeil,b Julia E. Ricea and James L. Hedrick*a
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

We report the effective organocatalysis of the aminolytic depolymerization of waste poly(ethylene


terephthalate) (PET) using 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) producing a broad range of
crystalline terephthalamides. This diverse set of monomers possesses great potential as building blocks
for high performance materials with desirable thermal and mechanical properties deriving from the
Received 25th September 2012
Accepted 20th November 2012
terephthalic moiety and amide hydrogen bonding. Further, a computational study established
mechanistic insight into self-catalyzed and organocatalyzed aminolysis of terephthalic esters, suggesting
DOI: 10.1039/c2py20793a
that the bifunctionality of TBD particularly concerning activation of the carbonyl group differentiates
www.rsc.org/polymers TBD from other organic bases.

Introduction However, it is still uncertain whether these advances have


enabled the establishment of a sustainable recycling system.
Recent advances in biomass transformation have enabled Chemical recycling is an ideal way of recovering pure mate-
poly(ethylene terephthalate) (PET), a widely used thermoplastic, rial.5,11 However, in comparison with mechanical recycling,
to be regarded as a bio-based polymer.1 The escalating chemical recycling will remain uncompetitive unless one
production of PET today is, however, still leading to a global develops innovative techniques to reduce production costs and
concern over the treatment of PET waste products and their energy consumption12 by reproducing the production of bottle-
destructive impact to the environment.2 Commonly practiced grade polymers in a highly efficient way. Nonetheless chemical
mechanical recycling has been considered as a temporary recycling can be leveraged as a standby for mechanical recycling
solution because the recycled PET, which is oen used for when polymers that have been degraded through repeated
secondary materials other than beverage bottles, ends up in mechanical recycling are chemically broken down to the
landlls. Early issues of the mechanical recycling were that the monomers.
residual metal catalysts used in the production of PET deterio- In addition, waste PET can be chemically processed to form
rate the physical properties of PET such as intrinsic viscosity monomers as feedstock for a variety of valued materials.13 In
and clarity during the material post-production process,2–5 and particular, the aminolysis of PET, which is more thermody-
that the bottle material contains not only PET homopolymer but namically favorable than alcoholysis, can potentially provide a
also copolymers derived from cyclohexanedimethanol and iso- spectrum of terephthalamides that can be broadly applicable to
phthalic acid3,5 and other polymers such as polyvinyl chloride the formation of additives, modiers, and building blocks for
and polyethylene-vinylacetate.6 Several innovations such as high performance materials.14–16 Aromatic amides used in these
solid-state polymerization,7 the addition of additives such as applications oen have enhanced mechanical and thermal
chain extenders,8 the formation of composites with virgin PET,9 properties due to hydrogen bonding and rigidity of struc-
and so-called super cleaning10 have enabled the establishment ture.16,17 Most terephthalamides are currently prepared through
of a bottle-to-bottle recycling process via mechanical recycling. amidation of dimethylterephthalate (DMT)18 or terephthaloyl
chloride (TC)19 monomers of PET. Several groups have reported
aminolytic depolymerization16,20–22 and modication of PET23
a
IBM Almaden Research Center, 650 Harry Road, San Jose, CA 95120, USA. E-mail: illustrating that generating useful materials from waste can be
fukushima@yz.yamagata-u.ac.jp; hedrick@us.ibm.com produced that could potentially alleviate existing environmental
b
Chemical and Materials Engineering Department, San Jose State University, San Jose, concerns and positively impact the management of limited
CA 95192, USA petroleum sources.
c
Physics and Astronomy Department, University of Denver, Denver, CO 80208, USA Although aminolysis reactions may be thermodynamically
d
King Abdul Aziz City for Science and Technology, Saudi Arabia
more favourable chemical transformations than alcoholysis,
† Electronic supplementary information (ESI) available: Experimental details and
the use of catalysts21 or microwave irradiation22 are still needed
Cartesian coordinates for optimized structures. See DOI: 10.1039/c2py20793a

1610 | Polym. Chem., 2013, 4, 1610–1616 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Polymer Chemistry

for efficient depolymerization. We recently reported that a


potent organic catalyst, 1,5,7-triazabicyclo[4.4.0]dec-5-ene
(TBD), positively impacts the depolymerization of PET by
ethylene glycol (EG), comparable to conventional metal cata-
lysts.24 Moreover, a recent report has demonstrated that TBD is
effective in aminolysis of various esters carried out at room
temperature under solventless conditions, indicating the
potency and versatility of the catalyst.25 These studies motivated
us to extend our organocatalytic depolymerization technique to
the aminolysis of PET thereby exploring alternative approaches
to the synthesis of terephthalamide compounds other than
those derived from DMT, and TC. Thus we investigated the
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

Fig. 1 General scheme of (a) the TBD-catalyzed aminolysis of PET and conver-
usefulness of TBD in the aminolytic depolymerization of PET
sions of PET into (b) 2a and (c) 2b as a function of reaction time. The aminolysis of
and surveyed the transformation of diverse amine compounds PET was conducted using 0.48 g of PET and 6 eq. of amine upon heating at 120  C
into a variety of terephthalamide compounds. for 2a and 180  C for 2b.
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

Experimental
specic time points. Since two amines are consumed to form a
Materials
terephthalamide, the end of the reaction (100% conversion)
Post-consumer PET beverage bottles were used as the PET was dened as when the ratio (terephthalamide : amine) is
source. The bottles were washed with water, dried in air at room 2 : (x  2), where x is the eq. amount of the amine reagents
temperature, and shredded by hand to a size of about 3 to 5 initially added. As a general rule, an excess of amine is used in
square mm. The PET akes were then further dried at 80  C these reactions since PET akes are amorphous and bulkier
under the vacuum overnight prior to use. TBD, amine reagents, than powdered PET or PET pellets. Thus, about 6 eq. of amine
and solvents were used as received (Sigma Aldrich). was used in reactions involving the depolymerization of PET
with benzyl amine (1a) and aniline (1b) (to be precise, 3.3 eq.
Measurements were used in the reaction involving 1a, and 3.0 eq. were used in
1 the reaction involving 1b in w/w). The temperatures at which the
H and 13C NMR spectra were obtained on a Bruker Avance 400
reactions were conducted were dependent on the reactivity and
Instrument at 400 MHz. Differential scanning calorimetry (DSC)
boiling or melting point of the reagent. As previously repor-
was recorded on a TA Instruments DSC Q2000 with a ramp rate
ted,20,21 the reactivity of amines in amidation reactions with
of 5  C min1 under a nitrogen atmosphere. A melting point of
esters are predominantly dictated by their basicity. The ami-
the product was determined as the onset temperature of the
nolysis of PET with 1a (pKa ¼ 9.34 (ref. 26a)) at 120  C was
melting endotherm. Thermal gravimetric analysis (TGA) was
complete in an hour in the presence of 5% TBD. When the
recorded on a TA Instruments TGA Q500 with a ramp rate of
reaction was run without catalyst, approximately 30% of PET
5  C min1 under a nitrogen purge.
was converted into 2a aer an hour and was still not complete
even aer 2 hours under these conditions (Fig. 1b). The ami-
General procedure for the TBD-catalyzed aminolysis of PET nolysis of PET using aniline (pKa ¼ 4.58 (ref. 26b)) proceeded
PET akes (0.48 g) were added into a 25 ml Schlenk tube con- quite slowly even though the reaction was performed at 180  C.
taining a reagent in excess (greater than 3 eq.) and 0.05 eq. of The reaction reached complete conversion aer 24 hours in the
TBD (17.5 mg) relative to a molar amount of the repeating unit presence of 5% TBD, but was only 50% complete in the absence
of PET (192). The tube was heated at a predetermined temper- of TBD over the same period of time, reecting the moderate
ature under nitrogen atmosphere in the absence of solvents. basicity of aniline (Fig. 1c). The discontinuous functional
Aliquots were taken for NMR analysis to monitor the conversion prole shown in Fig. 1c indicates that the reaction proceeds via
and determine the end of the reaction. Upon completion, the a multistep pathway.
product was isolated by trituration and washing with Depolymerization of PET is generally considered as a rst
organic solvents. Detailed procedures for the formation of each order reaction27 in cases involving the use of excess reagent
terephthalamide is provided in the ESI.† (greater than 10 eq.) as solvent. We have performed kinetic
studies of reactions involving 1a and 1b under second order
Results and discussion conditions because changes in concentration of the reagents
are not negligible. The semilogarithmic plots of time against
We rst estimated the catalytic activity of TBD in the aminolysis the PET conversion showed a linear correlation only in
of PET by determining the conversion of PET into the products the reaction involving 1a in the presence of 5% TBD (ESI,
as a function of reaction time using two aromatic amines with Fig. S1†), where the rate of reaction, k, was determined to be
or without catalyst (Fig. 1). Conversions were calculated as a 2.02  104 mol1 s1.
function of the molar ratio of the terephthalamide and amine Since the aminolysis of PET starts off as a nonhomogeneous
reagents as determined by integration of their NMR signals at system, the apparent reaction rate is dominated by mass

This journal is ª The Royal Society of Chemistry 2013 Polym. Chem., 2013, 4, 1610–1616 | 1611
View Article Online

Polymer Chemistry Paper

transfer and dissolution of PET into (melted) amine reagents, of bis-amine functionalized terephthalamides (2c–2k), because
rather than by the chemical reaction.23 In our studies, PET diamine compounds are in demand as monomers for highly
akes were not completely consumed during the induction thermostable polymers such as polyamides, polyimides, poly-
period, during which time the reaction rate hardly increases as urethanes, and polyureas, and also as a hardener/modier for
seen in Fig. S1.† Consequently, the kinetics could not be simply epoxy resins. To the best of our knowledge, only compounds 2c
evaluated by measuring concentration proles. These results and 2l have previously been synthesized through aminolysis of
suggest that TBD catalysis accelerates the mass transfer, leading PET.20–22 Unlike glycolysis,24 visual inspection was insufficient
to reduction of the induction period. Comparison of rate to determine the end of the aminolysis reactions. The products
constants aer the initial induction period, suggests that TBD crystallize out as the reactions proceed, and some reactions
promotes the reaction about ten times faster for the reaction to require further time for the completion aer the reaction
form 2a. The semilogarithmic plots for the reaction to form 2b, mixture becomes homogeneous. Thus, reaction completion
however, show exponential increase aer the induction period, was monitored by 1H NMR spectroscopy. Table 1 summarizes
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

implying that another concentration factor is involved. As the results of the TBD-catalyzed aminolysis reactions of PET
previously reported,24 ethylene glycol (EG) also serves as a using various amines. A small amount of oligomeric by-
catalyst. Thus, we speculate that the catalysis may be bolstered product was formed when using diamine reagents, most likely
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

by EG gradually formed from PET during the reaction, espe- as a dimer or trimer; the lone exception was observed for the
cially in reactions involving weakly basic amines such as 1b. reaction involving 4-aminobenzylamine (1k). Similar to
We then focused on producing functional terephthalamides glycolysis,24 the oligomeric by-product was reduced to a
as a potential key building block of high performance mate- minimal level by increasing reagent loading to 16 eq. (5 to
rials using different amine compounds. A diverse library of 11 eq. in w/w) relative to PET. As previously reported,24 the
terephthalamides was successfully developed from post- catalyst and excess reagent can also be recycled in the ami-
consumer PET via TBD-catalyzed aminolysis, as shown in nolysis to avoid wasting expensive reagents, although this
Fig. 2. Here, most of our interest was devoted to the formation resulted in longer reaction times.

Fig. 2 Terephthalamides produced through organocatalytic aminolysis of PET. See Table 1 for reaction conditions.

1612 | Polym. Chem., 2013, 4, 1610–1616 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Polymer Chemistry

Table 1 Summary of TBD-catalyzed aminolysis of PET using various amine reagentsa

Amineb eq. in mol


(eq. in w/w) pKac mpd (bp) ( C) Rxn temp. ( C) Rxn time (h) Yield (%) mp ( C)

1c 16 (5.0) 9.98 (ref. 26a) (118) 110 1 89 191


1d 16 (7.4) 10.65 (ref. 26d) 25–28 110 1 82 217
1e 16 (9.7) 11.02 (ref. 26c) 42–45 110 1 63 171
1f 16 (11.4) 9.46 (ref. 26c) 60–63 150 1 79 223
1g 16 (11.4) 9.19 (ref. 26e) (265) 150 1 67 128
1h 16 (9.1) 6.08 (ref. 26b) 138–143 190 18 72 301
1i 16 (9.1) 4.88 (ref. 26b) 64–66 190 18 75 295
1j 16 (9.2) 6.48 (ref. 26f) 117–122 190 18 72 240
1k 3 (2.0) 9.70 (ref. 26g) 37 120 2 70 203
1l 6 (2.1) 9.50 (ref. 26a) (170) 120 2 93 233
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

1m 3 (1.7) 9.91 (ref. 26h) (133) 120 2 74 173


1n 6 (1.9) 9.49 (ref. 26a) (53) 45 4 81 220
1o 22 (6.4) 8.15 (ref. 26c) (83) 70 135 66 205
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

a
PET (0.48 g) and TBD (17.5 mg) was used. b Relative to PET. c Ref. 26; measured at 25  C in water. d Melting points of reagents (boiling points are
shown in parentheses, where applicable).

Reactions involving alkyl and benzylic diamines were those involving primary amines because of steric congestion
complete in 1 to 2 hours, while reactions involving aryl around the nitrogen atom.
diamines 1h–1j, with moderate pKa's, were only complete aer With regards to 1g and 1i in Table 1, it is obvious that the
leaving them overnight, even upon heating at 190  C. Compared reactions would take longer time to recover products in around
to reactions involving 1c–1e, those involving 1h–1j require 80% yield if the reactions were conducted at 110  C where both
longer reaction times and higher temperatures for complete 1g and 1i have been completely liqueed. This indirectly
conversion. suggests that the basicity of amine reagents inuences the
However, similar to the reaction involving 1b, when per- efficiency of the aminolysis reaction.
formed in the absence of TBD the time required for completion The crude reaction mixture typically comprises the target
almost doubles. Reactions involving 1k and 1l provided terephthalamide, oligomeric by-product, excess reagent, and
mixtures of products at temperatures higher than 150  C. minimal catalyst and EG, which are easily separated. In the
Nevertheless, 2k and 2l were selectively formed via aminolysis at aminolysis of PET to form 2c conducted at 110  C for 2 h using
lower temperatures without undesired side products. The 16 eq. of ethylenediamine (1c, EDA), the weight ratios of iso-
formation of 2l by the aminolysis of PET has previously been lated 2c and the by-products were 50 : 6 (89% yield of tereph-
described.20–22 In terms of reaction time, reaction temperature, thalimide) for the reaction using TBD, 41 : 12 (77% yield of
and product yield, the TBD-catalyzed aminolysis for 2l reported terephthalimide) for the reaction using 1,8-diazabicyclo[5.4.0]
in this study was comparable to, or more favourable than, undec-7-ene (DBU), and 38 : 13 (74% yield of terephthalimide)
previous reports. Terephthalamides derived from amines 1n for the reaction using no catalyst, in accordance with the
and 1o, could also be formed but at longer reaction times due to observation that aminolytic depolymerizations of PET can be
their lower boiling points and reactivities. All products are completed in reasonable time without a catalyst in some cases
crystalline and most of the products were easily isolated by (e.g. EDA). The basicity of DBU (pKa ¼ 24.34 in acetonitrile)28 is
trituration and solid–liquid rinse with ordinary organic solvents similar to that of TBD (pKa ¼ 26.03 in acetonitrile),28 but DBU is
providing highly pure products with yields over 70%. Most of not as effective a catalyst in the reaction involving 1c, we
the products possessed melting points over 200  C. It appears surmised that bifunctional activation by TBD must play a
that the combination of the terephthalic moiety and hydrogen crucial role in the aminolysis.
bonding attributed to the amide fragments contribute to the To gain an understanding of the mechanism of aminolysis of
high melting points. The high yields obtained for these very PET with EDA in the absence of a catalyst, we performed density
pure compounds were controlled by the use of excess amine functional calculations with B3LYP/aug-cc-pVTZ//B3LYP/6-
reagent as well as active promotion of the reaction by TBD. 31+G* 29 in a continuum dielectric using IEF-cPCM30 as imple-
Chemoselectivity was achieved at low temperatures when mented in GAMESS-US.31 We chose an implicit solvent repre-
asymmetric amines were used. sentation of EDA with a dielectric of 8.9 to approximate
In addition to these reactions highlighting the TBD-catalyzed the experimental conditions of the PET aminolysis (in EDA at
aminolysis of PET with primary amines we have also performed 115  C).32
one example using a secondary amine. Piperidine (1p, pKa ¼ The self-catalysis model of aminolysis of esters was moti-
11.22)26a successfully depolymerized PET upon heating at 100  C vated by Schaefer's study on the ammonolysis of methylformate
for 16 h, providing the corresponding terephthalamide (2p) in (MF) in which two ammonia molecules form a reactant–catalyst
high yield (72%, see ESI†). Despite the strong basicity of dimer complex (Fig. 3a)33 (see ESI†). Thus, we investigated the
piperidine, this reaction is presumably slower in comparison to aminolysis of dimethylterephthalate (DMT, model for the

This journal is ª The Royal Society of Chemistry 2013 Polym. Chem., 2013, 4, 1610–1616 | 1613
View Article Online

Polymer Chemistry Paper

reactions involving DMT are 6–8 kcal mol1 higher. This is in


accordance with the observation that aromatic ester bonds are
harder to break than aliphatic ester bonds.24 Thus, the self-
catalyzed aminolysis of PET with EDA is plausibly explained by
the use of a large excess of EDA even though the barriers are
high in comparison with those observed for the ammonolysis of
MF and the glycolysis of PET in the presence of TBD.24,32
Fig. 3 Comparison of models of (a) two ammonia molecules during self-catalysis
A detailed computational investigation on the organo-
(N–N distance ¼ 3.2 Å) and (b) an EDA molecule (N–N distance ¼ 3.0 Å).
catalytic aminolysis of PET (repeat unit modeled by methyl-
benzoate) using TBD or DBU as catalyst is reported elsewhere.34
These results are discussed here for comparison and contrast
repeat unit of PET) with EDA (Fig. 4a). The barrier heights for with the data presented herein for the self-catalysis of the
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

reactions involving 1, 2 and 3 molecule(s) of EDA (1 EDA, 2 EDA aminolysis with EDA. The barrier heights for reactions involving
and 3 EDA) were 36.5 kcal mol1, 28.7 kcal mol1 and 27.3 kcal TBD were 22.9/20.9 (1 EDA/2 EDA) kcal mol1 and 25.9 (2 EDA)
mol1, respectively (Fig. 4b); as expected, the barrier decreases for reactions involving DBU. Despite the fact that DBU has only
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

when the number of EDA molecules increases. In comparison to one hydrogen bond acceptor, it is a stronger base, and thus a
the reaction involving the self-catalyzed aminolysis of MF (see better catalyst, than EDA itself. However, activation of the
ESI†), the reactivity trend is similar but, the barriers for carbonyl group by hydrogen bonding with TBD or EDA is an

Fig. 4 Scheme of aminolysis of dimethylterephthalate (DMT) using (a) EDA and (b) comparison of the rate limiting step in 1 EDA, 2 EDA, and (2 + 1) EDA pathways for
the aminolysis of DMT with EDA.

1614 | Polym. Chem., 2013, 4, 1610–1616 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Polymer Chemistry

important factor in the aminolysis of PET. The difference in the 12 (a) D. Carta, G. Cao and C. D'Angeli, Environ. Sci. Pollut. Res.,
computed barrier heights corresponds to the experimentally 2003, 10, 390–394; (b) H. Kurokawa, M. Oshima, K. Sugiyama
observed product weight ratios. and H. Miura, Polym. Degrad. Stab., 2003, 79, 529–533; (c)
In conclusion, we have demonstrated the effective organo- V. A. Kosmidis, D. S. Achilias and G. P. Karayannidis,
catalysis to promote aminolytic depolymerization of waste PET Macromol. Mater. Eng., 2001, 286, 640–647; (d)
using TBD as well as the production of a broad range of ter- M. S. Farahat and D. E. Nikles, Macromol. Mater. Eng.,
ephthalamide compounds that are crystalline with desirable 2001, 286, 695–704; (e) N. D. Pingale, V. S. Palekar and
thermal and mechanical properties. This diverse set of mono- S. R. Shukla, J. Appl. Polym. Sci., 2010, 115, 249–254.
mers shows a large potential as a building block of high 13 (a) S. R. Shulka, A. M. Harad and L. S. Jawale, Polym. Degrad.
performance materials due to their physical properties arising Stab., 2009, 94, 604–609; (b) M. Lu and S. Kim, J. Appl. Polym.
from terephthalic moiety and amide hydrogen bonding. The Sci., 2001, 80, 1052–1057; (c) K. Ertas and G. Güçlü, Polym.-
computational study ascertained the mechanistic insight of Plast. Technol. Eng., 2005, 44, 783–794; (d) C. Kawamura,
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

self-catalyzed aminolysis and organocatalyzed aminolysis of K. Ito, R. Nishida, I. Yoshihara and N. Numa, Prog. Org.
terephthalic esters, suggesting that the bifunctionality of TBD Coat., 2002, 45, 185–191; (e) D. E. Nikles and M. S. Farahat,
in particular for activation of the carbonyl group differentiates Macromol. Mater. Eng., 2005, 290, 13–30; (f) P. Radenkov,
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

TBD from other organic bases. M. Radenkov, G. Grancharov and K. Troev, Eur. Polym. J.,
2003, 39, 1223–1228.
Notes and references 14 (a) M. Van der Schuur, B. Noordover and R. J. Gaymans,
Polymer, 2006, 47, 1091–1100; (b) W. Fischer and
1 (a) Y. Román-Leshkov, J. N. Chheda and J. A. Dumesic, C. Helbling, EP 19940810643, 1995.
Science, 2006, 312, 1933–1937; (b) R. M. Kriegel, X. Huang 15 (a) T. J. Stierman, US 5212261, 1993; (b) G. Raspanti and
and M. W. Schultheis, US 20090246430, 2009. M. Brena, EP 0575836, 1997; (c) R. V. Sparer and
2 (a) L. M. Rios, C. Moore and P. R. Jones, Mar. Pollut. Bull., S. A. Pogany, US 4549010, 1985.
2007, 54, 1230–1237; (b) J. Scheirs, Polymer Recycling: Science, 16 T. Spychaj, E. Fabrycy, S. Spychaj and M. Kacperski, J. Mater.
Technology and Applications, Wiley, New York, 1998, ch. 4. Cycles Waste Manage., 2001, 3, 24–31.
3 D. D. Cornell, in Modern Polyesters: Chemistry and Technology 17 (a) J. P. Sheth, D. B. Klinedinst, T. W. Pechar, G. L. Wilkes,
of Polyesters and Copolyesters, ed. J. Scheirs and T. E. Long, E. Yilgor and I. Yilgor, Macromolecules, 2005, 38, 10074–
Wiley, New York, 2003, ch. 16, pp. 565–587. 10079; (b) E. J. Woo, G. Farber, R. J. Farris, C. P. Lillya and
4 H. Zimmerman and N. T. Kim, Polym. Eng. Sci., 1980, 20, J. C. W. Chien, Polym. Eng. Sci., 1985, 25, 834–840.
680–683. 18 (a) J. Krijgsman, D. Husken and R. J. Gaymans, Polymer,
5 P. Klein, in Recycling and Recovery of Plastics, ed. J. Brandrup, 2003, 44, 7043–7053; (b) L. Nery, H. Lefebvre and A. Fradet,
M. Bittner, G. Menges and W. Michaeli, Hanser Gardner Macromol. Chem. Phys., 2003, 204, 1755–1764.
Publ., Cincinatti OH, 1996, ch. 5, pp. 494–502. 19 (a) H. Li, C. Liu, G. Guo, X. Zhou, J. Zhang, D. Shen, Z. Zhang,
6 (a) N. Torres, J. J. Robin and B. Boutevin, Eur. Polym. J., 2000, P. Xie, S. Yu and R. Zhang, J. Polym. Sci., Part A: Polym. Chem.,
36, 2075–2080; (b) A. Pawlak, M. Pluta, J. Morawiec, 2002, 40, 3161–3170; (b) I. Aoki and S. Shinkai, JP 06192224,
A. Galeski and M. Pracella, Eur. Polym. J., 2000, 36, 1875– 1994; (c) G. A. Sotzing, J. R. Reynolds, A. R. Katritzky,
1884; (c) M. Paci and F. P. L. Manita, Polym. Degrad. Stab., J. Soloducho, S. Belyakov and R. Musgrave,
1998, 61, 417–420. Macromolecules, 1996, 29, 1679–1684.
7 (a) H. A. Ghatta, S. Cobor and T. Severini, Polym. Adv. 20 H. Zahn and H. Pfeifer, Polymer, 1963, 4, 429–452.
Technol., 1997, 8, 161–168; (b) I. Kenoka and Y. Kawakami, 21 (a) S. R. Shukla and A. M. Harad, Polym. Degrad. Stab., 2006,
JP 2000219728, 2000. 91, 1850–1854; (b) M. E. Tawk and S. B. Eskander, Polym.
8 (a) S. Japon, L. Boogh, Y. Leterrier and J. E. Manson, Polymer, Degrad. Stab., 2010, 95, 187–194.
2000, 41, 5809–5818; (b) N. Torres, J. J. Robin and 22 N. D. Pingale and S. R. Shukla, Eur. Polym. J., 2009, 45, 2695–
B. Boutevin, J. Appl. Polym. Sci., 2001, 79, 1816–1824; (c) 2700.
X. Xua, Y. Dinga, Z. Qiana, F. Wanga, B. Wena, H. Zhoua, 23 (a) M. Avadanei, M. Drobota, I. Stoica, E. Rusu and
S. Zhanga and M. Yang, Polym. Degrad. Stab., 2009, 94, V. Barboiu, J. Polym. Sci., Part A: Polym. Chem., 2010, 48,
113–123; (d) B. Jacques, J. Devaux, R. Legras and E. Nield, 5456–5467; (b) S. Nakano, US 6107438, 2000.
Polymer, 1996, 37, 1189–1200; (e) D. M. Conor and 24 K. Fukushima, O. Coulembier, J. M. Lecuyer, H. A. Almegren,
K. A. Keller, US 20060148914, 2006. A. M. Alabdulrahman, F. D. Alsewailem, M. A. McNeil,
9 F. M. Schloss and D. C. Balduff, US 5965081, 1999. P. Dubois, R. M. Waymouth, H. W. Horn, J. E. Rice and
10 (a) A. S. F. Santos, J. A. M. Agnelli and S. Manrich, Food Addit. J. L. Hedrick, J. Polym. Sci., Part A: Polym. Chem., 2011, 49,
Contam., Part A, 2010, 27, 567–573; (b) G. W. Bohnert and 1273–1281.
T. E. Hand and G. M. DeLaurentiis, US 7473758, 2009; (c) 25 C. Sabot, K. A. Kumar, S. Meunier and C. Mioskowski,
M. Gupta, S. A. Bandi, S. Mehta and D. A. Schiraldi, J. Appl. Tetrahedron Lett., 2007, 48, 3863–3866.
Polym. Sci., 2008, 107, 3212–3220. 26 (a) H. K. Hall Jr, J. Am. Chem. Soc., 1957, 79, 5441–5443; (b)
11 D. Paszun and T. Spychaj, Ind. Eng. Chem. Res., 1997, 36, K. C. Gross and P. G. Seybold, Int. J. Quantum Chem., 2000,
1373–1383. 80, 1107–1115; (c) D. D. Perrin, Dissociation Constants of

This journal is ª The Royal Society of Chemistry 2013 Polym. Chem., 2013, 4, 1610–1616 | 1615
View Article Online

Polymer Chemistry Paper

Organic Bases in Aqueous Solution, Butterworths, London, G. W. Spitznagel and P. v. R. Schleyer, J. Comput. Chem.,
1965; Supplement, 1972; (d) A. E. Martell and R. M. Smith, 1983, 4, 294–301; (g) D. E. Woon and T. H. Dunning,
Critical Stability Constants, Plenum Press, New York, 1974, J. Chem. Phys., 1995, 103, 4572–4585.
vol. 2; (e) B. Gawe1 and W. Lasocha, Z. Kristallogr. Suppl., 30 V. Barone and M. Cossi, J. Phys. Chem. A, 1998, 102, 1995–
2007, 26, 611–616; (f) X. Shi, I. Banyai, W. G. Lesniak, 2001.
M. T. Islam, I. Orszagh, P. Balogh, J. R. Baker, Jr and 31 M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert,
L. P. Balogh, Electrophoresis, 2005, 26, 2949–2959; (g) M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga,
J. F. King, R. Rathore, J. Y. L. Lam, Z. R. Guo and K. A. Nguyen, S. Su, T. L. Windus, M. Dupuis and
D. F. Klassen, J. Am. Chem. Soc., 1992, 114, 3028–3033; (h) J. A. Montgomery, J. Comput. Chem., 1993, 14, 1347–1363.
J. Hine, F. A. Via and J. H. Jensen, J. Org. Chem., 1971, 36, 32 Equations of state for 3(T) were obtained from CRC
2926–2929. Handbook of Chemistry and Physics, 84th edn., CRC
27 N. Halacheva and P. Novakov, Polymer, 1995, 36, 867–874. Press, Boca Raton, Fl. 2003, vol. 1, pp. 6–158 and
Published on 19 December 2012 on http://pubs.rsc.org | doi:10.1039/C2PY20793A

28 I. Kaljurand, A. Kutt, L. Soovali, T. Rodima, V. Maemets, extrapolated 3(1/T) to the temperature under which the
I. Leito and I. A. Koppel, J. Org. Chem., 2005, 70, 1019–1028. reactions are conducted experimentally.
29 (a) A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652; (b) 33 S. IIieva, B. Galabov, D. G. Musaev, K. Morokuma and
Downloaded by ST LAWRENCE UNIVERSITY on 12 February 2013

C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. H. F. Schaefer III, J. Org. Chem., 2003, 68, 1496–1502.
Matter Mater. Phys., 1988, 37, 785–789; (c) S. H. Vosko, 34 H. W. Horn, G. O. Jones, D. S. Wei, K. Fukushima,
L. Wilk and M. Nusair, Can. J. Phys., 1980, 58, 1200–1211; J. M. Lecuyer, J. L. Hedrick and J. E. Rice, Computational
(d) P. J. Stephens, F. J. Devlin, C. F. Chabalowski and investigations on organocatalytic amidation and trans-
M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627; (e) esterication of aromatic esters as a model for the
W. J. Hehre, R. J. Ditcheld and J. A. Pople, J. Chem. Phys., depolymerization of polyethylene terephthalate, J. Phys.
1972, 56, 2257–2261; (f) T. Clark, J. Chandrasekhar, Chem. A, 2012, DOI: 10.1021/jp304212y.

1616 | Polym. Chem., 2013, 4, 1610–1616 This journal is ª The Royal Society of Chemistry 2013

View publication stats

You might also like