You are on page 1of 10

Green Chemistry

View Article Online


PAPER View Journal

Facile synthesis of hemiacetal ester-based dynamic


covalent polymer networks combining fast
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

Cite this: DOI: 10.1039/d1gc02773e


reprocessability and high performance†
Hongzhi Feng,a,b Songqi Ma, *a Xiwei Xu,a,b Qiong Li,a,b Binbo Wang,a,b Na Lu,a,b
Pengyun Li,a,b Sheng Wang,a,b Zhen Yua and Jin Zhua

Dynamic covalent polymer networks (DCPNs) can address the recycle or reuse issue of thermosets due
to their network rearrangement from the dynamic bond exchange. However, developing DCPNs combin-
ing fast reprocessability and high performance through a facile and green method is still a huge challenge.
Herein, a facile and green method (in situ polymerization and dynamic cross-linking (ISPDC)) was utilized
to produce hemiacetal ester-based DCPNs which exhibited fast reprocessability and high performance.
No solvent and no purification were required, and the facile synthesis process also showed readily regu-
lated performance for the DCPNs. The obtained DCPNs could be continuously reprocessed by extrusion
as well as compression remolding; meanwhile their solvent resistance, thermo-physical and mechanical
properties are excellent on account of their high cross-link density. The small-molecule model reactions
Received 3rd August 2021, and theoretical calculation demonstrated that hemiacetal ester exchange follows a dissociative mecha-
Accepted 8th October 2021
nism without or with an extra carboxyl group. This work studied DCPNs which showed great potential to
DOI: 10.1039/d1gc02773e be efficiently recycled via reprocessing and scaled up. It tallies with sustainable development and contrib-
rsc.li/greenchem utes to carbon neutrality.

1. Introduction networks have attracted great attention. In virtue of the cross-


linked networks, they possess the high performance of thermo-
Thermosets are covalently cross-linked three-dimensional sets; and on account of the dynamic bonds, they can be repro-
polymer networks.1,2 Due to their superior thermo-physical cessed like thermoplastics and have functions such as self-
and mechanical properties, chemical resistance, and dimen- healing, degradability, recyclability, weldability, shape change-
sional stability from their network features, thermosets have ability, etc. So far, plenty of dynamic bonds or chemistries such
been widely used in coatings, adhesives, composites, elec- as ester bond,14 D–A adduct,15,16 disulfide bond,17,18 boronic
tronic packaging, etc., and are irreplaceable in many esters,19–22 olefin metathesis,23 Schiff base,24–26 vinylogous
applications.2–7 However, their permanently cross-linked net- urethanes/ureas,27–30 thiol–disulfide exchange,31 transalkyla-
works make thermosets arduous to be recycled or reprocessed tion,32 silyl ether,33,34 hemiaminals,35 thioacetal,36 hindered
after use, and the main disposal methods are landfilling and urea bonds,37 carbamates,38 diketoenamine,4 diselenide,39 and
incineration,7 which are against the sustainable development acetal40–43 have been explored to produce DCPNs. However, due
of materials. to the slow network rearrangement speed, the overwhelming
In recent decades, covalent adaptable networks (CANs)8,9 or majority of reported DCPNs require a long reprocessing time at
vitrimers10,11 or dynamic covalent polymer networks elevated temperatures, which not only could damage their net-
(DCPNs)12,13 which possess dynamic bonds in the cross-linked works and consequently decrease their properties, but also
could consume a lot of energy. Besides, they are limited to com-
pression remolding (incontinuous reprocessing method) and
a
Key Laboratory of Bio-based Polymeric Materials Technology and Application of not applicable to continuous reprocessing methods such as
Zhejiang Province, Laboratory of Polymers and Composites, Ningbo Institute of extrusion and injection.
Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo
Some efforts have been made to accelerate the dynamic
315201, P. R. China. E-mail: masongqi@nimte.ac.cn; Tel: +86-574-87619806
b
University of Chinese Academy of Sciences, Beijing 100049, P. R. China
exchange of the DCPNs and a few continuously reprocessable
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ DCPNs have been achieved.14,28,44–47 The first work was from
d1gc02773e Leibler and coworkers,14 and they introduced 5 mol% Zn(Ac)2

This journal is © The Royal Society of Chemistry 2021 Green Chem.


View Article Online

Paper Green Chemistry

(as the catalyst for transesterification) into a soft epoxy-acid and propionic acid (PPA) ( purity ≥99%) were purchased from
network and acquired injection remolding ability. Winne and Macklin Reagent Co., Ltd. 1,4-Cyclohexanedimethanol divinyl
Du Prez and coworkers28 also introduced catalyst pTsOH ether (CDVE, a mixture of the cis/trans isomers) was obtained
(6 mol%) and pendent amino groups and designed a favorable from Sigma-Aldrich Co., China. Isobutyl vinyl ether (IVE)
macromolecular architecture for vinylogous urethane vitrimers ( purity ≥99.5%), cyclohexyl vinyl ether (CVE) (a mixture of the
to gain extrusion capacity. The double neighboring group par- cis/trans isomers, purity ≥99%), benzoyl peroxide (BPO) (AR)
ticipation method was also disclosed to accelerate the dynamic and anhydrous magnesium sulfate (MgSO4) (AR) were pur-
exchange and network rearrangement of phthalate monoester- chased from Aladdin Reagent Co., China. Methylbenzene, di-
based DCPNs, and extrudable DCPNs were obtained.44 chloromethane (CH2Cl2), methanol, ethanol and sodium car-
Commercial thermoplastics such as PE and PET together with bonate were bought from Sinopharm Chemical Reagent Co.,
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

dynamic cross-linking techniques were also utilized to Ltd, China. Anhydrous solvents were dried using an activated
produce continuously reprocessable DCPNs.11,46,47 Although molecular sieve before being used. All other chemicals were
significant advances have been achieved, developing continu- used as received unless noted.
ously reprocessable DCPNs combining fast reprocessability
and high performance via a facile and green method is still a
huge challenge.
2.2. Preparation of hemiacetal ester dynamic covalent
Recently, Ladmiral and coworkers48 for the first time
polymer networks
reported the catalyst-free dynamic exchange reaction of a car-
boxyl group with hemiacetal ester and utilized it to produce A series of hemiacetal ester dynamic covalent polymer net-
polystyrene-based vitrimers which could be reprocessed works were prepared, and the total molar amount of MMA and
between 80 and 130 °C. In this work, we further exploited the MAA in each sample was 0.2 mol. The amount of initiator
dynamic feature of hemiacetal ester and utilized it to achieve (BPO) added was 1 mol% of the total molar mass of MAA and
continuously reprocessable DCPNs with high performance. In MMA. In a representative procedure for the preparation of
addition, a facile, green and efficient synthetic method—in situ PMMC-10, MMA (18.0216 g, 0.18 mol), MAA (1.7218 g,
polymerization and dynamic cross-linking (ISPDC)—was devel- 0.02 mol), CDVE (1.9629 g, 0.01 mol) and BPO (0.4866 g,
oped to produce hemiacetal ester DCPNs. In this process, all 2 mmol) were mixed and poured into a 100 mL round-bot-
the raw materials, methacrylic acid, methyl methacrylate and tomed flask with a magnetic stirrer. The mixture was stirred in
1,4-cyclohexanedimethanol divinyl ether, were commercially a water bath at 85 °C for about 10 min until it had a certain vis-
available, no solvent and no purification were required, and a cosity (similar to glycerol); at this point the conversion rate is
high atom conversion rate was achieved, which showed great about 10%, then the mixture was cooled quickly with ice water
economic and ecological advantages. The dynamic exchange to room temperature. The pre-polymerized material was
mechanism of hemiacetal ester linkage was investigated injected into a glass plate mould made in-house, which was
through small-molecule model reactions and theoretical calcu- polymerized at 60 °C for 2 h, 80 °C for 8 h, and finally treated
lation. The thermo-physical and mechanical properties were at 120 °C for 2 h. The synthetic route is illustrated in
evaluated by examining the glass transition temperature and Scheme 1. Other formulations of hemiacetal ester dynamic
tensile properties of the DCPNs. The malleability and the covalent polymer networks are summarized in Table 1.
reprocessability of the DCPNs were studied via stress relaxation
examination, compression remolding (incontinuous), and
extrusion (continuous). Table 1 Feed compositions of the hemiacetal ester covalent adaptable
networks

Compositions (mol%)

2. Experimental section Sample BPO MMA MAA CDVE


2.1. Materials PMMC-5 1 95 5 2.5
PMMC-10 1 90 10 5
Methacrylic acid (MAA) ( purity >99.0%), methyl methacrylate PMMC-15 1 85 15 7.5
(MMA) ( purity >99.5%), isobutyric acid (IBA) ( purity ≥99%), PMMC-20 1 80 20 10

Scheme 1 Synthetic route of the hemiacetal ester dynamic covalent polymer networks.

Green Chem. This journal is © The Royal Society of Chemistry 2021


View Article Online

Green Chemistry Paper

2.3. Synthesis of small-molecule model hemiacetal esters Isobutyric acid–isobutyl vinyl ether hemiacetal ester (IBA–
The small-molecule model isobutyric acid–cyclohexyl vinyl IVE): 1H NMR (400 MHz, DMSO-d6) δ 5.83 (q, J = 5.2 Hz, 1H),
ether hemiacetal ester (IBA–CVE) was synthesized from isobu- 3.27 (m, J = 40.0, 9.3, 6.6 Hz, 2H), 2.58–2.46 (m, 2H), 1.75 (m, J
tyric acid and cyclohexyl vinyl ether. Isobutyric acid (4.41 g, = 13.3, 6.7 Hz, 1H), 1.31 (d, J = 5.2 Hz, 3H), 1.09 (dd, J = 8.7,
0.05 mol) and cyclohexyl vinyl ether (6.31 g, 0.05 mol) were 7.0 Hz, 6H), 0.85 (d, J = 6.7 Hz, 6H).
mixed without any solvent and stirred at 80 °C for 3 h. Then,
13
C NMR (151 MHz, DMSO-d6) δ 176.16, 96.14, 75.13, 33.90,
the unreacted isobutyric acid and cyclohexyl vinyl ether were 28.33, 20.89, 19.46, 19.44, 19.20, 18.98.
removed by vacuum distillation. The residue was mixed with Propionic acid–cyclohexyl vinyl ether hemiacetal ester (PPA–
saturated Na2CO3 aqueous solution (50 mL), and extracted CVE): 1H NMR (400 MHz, DMSO-d6) δ 5.96 (q, J = 5.2 Hz, 1H),
3.52 (tt, J = 8.8, 4.0 Hz, 1H), 2.29 (qd, J = 7.5, 2.7 Hz, 2H), 1.03
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

with dichloromethane (50 mL). The organic phases were com-


bined, washed with deionized water (100 mL × 3), and dried (t, J = 7.5 Hz, 3H).
with anhydrous MgSO4. The solvent was removed on a rotary
13
C NMR (151 MHz, DMSO-d6) δ 173.65, 94.18, 75.77, 33.25,
evaporator to afford the model hemiacetal ester to obtain the 31.86, 27.58, 25.58, 23.88, 23.71, 21.61, 9.22.
model hemiacetal ester IBA–CVE. PPA–IVE from propionic acid
and isobutyl vinyl ether, IBA–IVE from isobutyric acid and iso- 2.4. Gel content test
butyl vinyl ether, PPA–CVE from propionic acid and cyclohexyl The samples (including PMMC-5, PMMC-10, PMMC-15 and
vinyl ether were synthesized by the same method. The syn- PMMC-20) (around 300 mg) were separately put into a Soxhlet
thetic routes are illustrated in Scheme 2. extractor, and extracted with methylbenzene for 48 h, then
Isobutyric acid–cyclohexyl vinyl ether hemiacetal ester (IBA– dried at 85 °C in a vacuum oven for 24 h. m0 is the initial
CVE): 1H NMR (400 MHz, DMSO-d6) δ 5.95 (q, J = 5.2 Hz, 1H), mass, and m1 is the final mass after drying; the gel content is
3.49 (dp, J = 8.8, 3.9 Hz, 1H), 2.50 ( p, J = 7.0 Hz, 1H), 1.82 ( p, J calculated by m1/m0 × 100%.
= 5.0 Hz, 1H), 1.76–1.67 (m, 1H), 1.70–1.60 (m, 2H), 1.46 (tt, J =
7.8, 2.8 Hz, 1H), 1.34–1.15 (m, 5H), 1.09 (t, J = 6.8 Hz, 5H). 2.5. Swelling measurement
13
C NMR (151 MHz, DMSO-d6) δ 175.73, 93.98, 75.49, 33.49, To prove cross-linking, approximately 50 mg of PMMC-5 was
32.86, 31.46, 25.11, 23.46, 23.30, 21.16, 18.69, 18.51. immersed in sufficient methylbenzene (about 15 mL) at room
Propionic acid–isobutyl vinyl ether hemiacetal ester (PPA– temperature for 48 h. Then the sample was taken out and the
IVE): 1H NMR (400 MHz, DMSO-d6) δ 5.83 (q, J = 5.2 Hz, 1H), solvent on the surface was wiped completely with absorbent
3.34 (dd, J = 9.2, 6.5 Hz, 1H), 3.21 (dd, J = 9.3, 6.6 Hz, 1H), 2.32 cotton. mi is the initial mass and mf is the final mass after the
(qd, J = 7.5, 5.3 Hz, 2H), 1.75 (m, J = 13.3, 6.7 Hz, 1H), 1.30 (d, J swelling experiment. The swelling ratio is calculated by 100%
= 5.2 Hz, 3H), 1.03 (t, J = 7.5 Hz, 3H), 0.85 (dd, J = 6.7, 1.0 Hz, × (mf − mi)/mi.
6H).
13
C NMR (151 MHz, DMSO-d6) δ 173.27, 95.59, 74.66, 27.82, 2.6. Rheological behavior
26.96, 20.49, 18.98, 8.83.
The rheological experiments were carried out on a Discovery
HR-3 Rheometer (TA Instrument, America) with a plate dia-
meter of 25 mm shear geometry. The measurements were per-
formed only once, and sometimes repeated a second time if
the results were irregular. An initial compression force of 5 N
was applied and then the measurements were performed at a
fixed gap and a suitable deformation within the linear visco-
elastic region. The samples had a thickness of about 5 mm
and were cut into wafers with a diameter of 25 mm.
Stress relaxation. Stress relaxation tests were carried via the
stress relaxation mode. In order to avoid the deformation of
the samples when heated to the test temperature, the samples
were preloaded with a force of 1 × 10−3 N. When the tempera-
ture rose to the test temperatures (160 °C, 170 °C, 180 °C and
190 °C), the samples were kept for 10 min to reach thermal
equilibrium. During the tests, a constant strain (5%) was
applied to each sample, and the relaxation modulus was
recorded at a constant temperature.
Calculation of activation energies (Eas). Activation energies
(Eas) were determined using the methodology reported in the
literature.10 The measured values of characteristic relaxation
Scheme 2 Synthetic routes of small-molecule model hemiacetal times (τ*s) were plotted versus 1000/T. The plots were fit to the
esters. Arrhenius law in eqn (1)

This journal is © The Royal Society of Chemistry 2021 Green Chem.


View Article Online

Paper Green Chemistry

τ*ðTÞ ¼ τ*0 eEa=RT ð1Þ 2.11. Theoretical calculation


−1 −1 The Gaussian 16 program was employed for DFT calculations.
R: universal gas constant; 8.314 J K mol .
Frequency sweep experiments. Frequency sweep experiments Geometry optimization was performed at the B3LYP-D3BJ/6-
were carried via the frequency sweep mode. The frequency was 31G(d,p) level,49–52 and frequency calculation was followed to
varied from 0.1 to 500 rad s−1 and performed in the tempera- ensure minimum or transition state was located, and to obtain
ture range between 140 and 190 °C with a deformation of 1%. Gibbs free energy, the SMD implicit solvation model53 was
used to take account of the solvation effect; although the reac-
2.7. Reprocess recycling through hot pressing tion was performed without extra solvent, the solvation effect
originating from the substrates is non-negligible, and ethyl
The reprocess recycling was carried out through a flat vulcani-
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

acetate was employed as the implicit solvation, in order to rep-


zer (ZG-60T, Dongguan ZhengGong Electromechanical
resent the polar effect of the substrate mixture.
Equipment Technology Co., Ltd). The samples (including
PMMC-5, PMMC-10, PMMC-15 and, PMMC-20) were cut into 2.12. Characterization
small pieces and placed between two steel sheets covered with
Nuclear magnetic resonance (NMR) spectroscopy was recorded
a PI ( polyimide) film. After preheating at 160 °C for 2 min, the
on a Bruker AVANCE II spectrometer at 400 MHz (1H NMR) and
pressure was raised to 10 MPa and maintained for a certain
Bruker AVANCE III spectrometer at 151 MHz (13C NMR), respect-
time (according to samples and experiments). After cooling to
ively, using DMSO-d6 as the solvent. Fourier-transform infrared
room temperature, reprocessed hemiacetal ester networks were
(FTIR) spectroscopy was recorded with a Micro-FTIR Cary660
obtained.
(Agilent, America) in the solid state and recorded in attenuated
total reflectance (ATR) mode, respectively. Differential scanning
2.8. Reprocess recycling through an extrusion plastometer
calorimetry (DSC) was examined via a Mettler-Toledo Star 1
Extrusion of the hemiacetal ester dynamic covalent polymer apparatus to investigate the glass transition temperatures of the
network was implemented via an extrusion plastometer hemiacetal ester networks. The cured samples (around 6 mg)
(YK-3651A, Dongguan YK Group). According to the standard were heated to 180 °C from 0 °C at a heating rate of 10 °C
test method for melt flow rates of thermoplastics using an min−1, and held at 180 °C for 3 min, cooled to 25 °C at a
extrusion plastometer (ASTM D-1238), about 5 g of PMMC-5 cooling rate of 20 °C min−1, and finally heated to 200 °C at a
which was crushed using a shredder was added to the cavity of heating rate of 10 °C min−1. The second heating curve was used
the plastometer, after keeping at 200 °C for 2 min, 3.18 kg was to determine glass transition temperatures for each sample.
applied to extrude the PMMC-5. Thermogravimetric analysis (TGA) was performed on a Mettler
Toledo TGA/DSC1 under a nitrogen atmosphere from 50 °C to
2.9. Reprocess recycling through a twin-screw mini-extruder 800 °C at a heating rate of 10 °C min−1 with around 6 mg of
Extrusion was performed on a twin-screw mini-extruder sample. Dynamic mechanical analysis (DMA) was used to
(WLG10/WZS10-D). The extruder was preheated at 200 °C for obtain the storage modulus and tan delta on a Q800 DMA (TA
10 min and fed with PMMC-5 powder at a rotational speed of Instruments, American). The test was performed in a tension
the screws of 15 rpm. The PMMC-5 powder above was dried mode at a frequency of 1 Hz. All the samples were cut into rec-
thoroughly under vacuum for 6 h at 80 °C. tangles with dimensions of around 25 mm (length) × 5 mm
(width) × 0.5 mm (thickness) to test the dynamic mechanical
2.10. Exchange reactions of small-molecule model properties of the hemiacetal ester networks from −25 to 200 °C
hemiacetal esters at a ramp rate of 3 °C min−1. GC-MS spectra were obtained
using a 7890B-5977A gas chromatography-mass spectrometer
In the acid-free system, small-molecule models PPA–IVE (GC-MS) (Agilent, America) using dichloromethane as the
(1.0234 g, 5 mmol) and IBA–CVE (1.2587 g, 5 mmol) were solvent. Tensile properties of the samples were studied on an
placed in a 10 mL vial, preheated in an oil bath at different Instron 5567 Electric Universal Testing Machine (Instron,
temperatures for 3 h, and 10 μL of each mixture was taken out America). The samples with dimensions of 40 mm (length) ×
and diluted in 990 μL of CHCl2 for gas chromatography-mass 5 mm (width) × 0.5 mm (thickness) were measured with a
spectrometry (GC-MS) examination. At 120 °C, 10 μL of the gauge length of 20 mm at a cross-head speed of 2 mm min−1.
mixture was taken out at different time durations and diluted For accuracy, the tensile properties of each sample were
in 990 μL of CHCl2 for GC-MS examination. In the acid system, reported as the average of five repeated measurements.
small-molecule models PPA–IVE (1.0234 g, 5 mmol), IBA–CVE
(1.2587 g, 5 mmol) and propionic acid (0.0370 g, 0.05 mmol)
were placed in a 10 mL vial, preheated in an oil bath at 3. Results and discussion
different temperature for 3 h, and 10 μL of each mixture was
taken out and diluted in 990 μL of CHCl2 for gas chromato- 3.1. Synthesis and chemical characterization of hemiacetal
graphy-mass spectrometry (GC-MS) examination. At 60 °C, ester DCPNs
10 μL of mixture was taken out at different time durations and The hemiacetal ester DCPNs, PMMC-X (X is the molar percen-
diluted in 990 μL of CHCl2 for GC-MS examination. tage (mol%) of the carboxyl group based on comonomers, X =

Green Chem. This journal is © The Royal Society of Chemistry 2021


View Article Online

Green Chemistry Paper

5, 10, 15, 20), were synthesized via a facile method of in situ 3.2. Thermo-physical and mechanical properties of the
polymerization and dynamic cross-linking (free radical hemiacetal ester DCPNs
copolymerization and the carboxyl–vinyl ether addition reac-
tion in one pot) (Fig. 1a, Scheme 1 and Table 1). The chemi- The PMMCs possessed favorable thermo-physical and mechan-
cal structures of the obtained hemiacetal ester DCPNs were ical properties, which greatly increased its application poten-
determined by FTIR (Fig. S1†). As shown in the FTIR spec- tial. Owing to the flexible hemiacetal ester linkage, the glass
trum of one typical hemiacetal ester DCPN PMMC-10 transition temperature (Tg) of the hemiacetal ester DCPNs
(Fig. 1b), the signals for carbon–carbon double bonds from decreased with increasing cross-linker CDVE content (corres-
the comonomers MMA and MAA and cross-linker CDVE at ponding to the content of hemiacetal ester linkage) from
around 1700 cm−1 disappeared. In addition, the successful 126 °C for PMMC-5 to 113 °C for PMMC-20 (Fig. 1e). For the
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

synthesis of the small-molecule hemiacetal ester model cross-linked polymers, the Tg has a close tie with cross-link
compounds (Scheme 2 and Fig. S2–S9†) demonstrated density and the rigidity of their chain segment.54,55 The cross-
the formation of hemiacetal ester linkage at the curing link density (ve) of the hemiacetal ester DCPNs were calculated
condition. These results are indicative of the free radical from the DMA date via eqn (2):
polymerization of MMA and MAA and the cross-linking reac-
E′ ¼ 3ve RT ð2Þ
tion between the carboxyl group from MAA and vinyl
ether group from CDVE. The gel contents of these DCPNs where E′ is the storage modulus of the PU networks at Tg +
were in the range of 93% to 96% (Fig. 1c) which are relatively 30 °C, R is the gas constant (8.314 J mol−1 K−1), and T is the
high, and the networks kept stable during the swelling kelvin temperature of Tg + 30 °C. As shown in Table S1,† the
experiments (Fig. 1d) with the swelling ratios about 52%, hemiacetal ester DCPNs have a lower cross-link density, which
which also indicates the successful synthesis of the DCPNs. might be due to the dissociation of the hemiacetal ester bond

Fig. 1 Preparation and properties of the hemiacetal ester DCPNs (PMMCs). (a) Schematic diagram of the preparation of the PMMCs. (b) FTIR spectra
of MMA, MAA, CVDE and PMMC-10. (c) Gel content of PMMA and PMMCs in toluene. (d) Digital pictures of the PMMC-10 before and after swelling
in toluene at room temperature for 48 h. (e) Storage modulus and tan δ as a function of temperature for the PMMCs by DMA. (f ) Representative
stress–strain curves of the PMMCs. (g) TGA curves of the PMMCs.

This journal is © The Royal Society of Chemistry 2021 Green Chem.


View Article Online

Paper Green Chemistry

at those temperatures. They showed superior mechanical pro- 3.3. Malleability and reprocessability of the hemiacetal ester
perties especially Young’s moduli (1773–1967 MPa) and tensile DCPNs
strength (53–57 MPa) which are similar to or even higher than
those of the commercial thermoplastic polymethyl methacry- The stress relaxation times (τ*s) of DCPNs can demonstrate
late (PMMA, Young’s modulus: 1760 MPa and tensile strength: their reprocessability. The PMMCs relaxed rapidly at high
39 MPa (ref. 11)). Similarly, their Young’s moduli reduced with temperatures (Fig. S10–S13†), and full relaxation required a
increasing cross-linker CDVE content, while there is not too short time. At 190 °C, they could relax to 1/e of the initial
much of a difference for their elongation at break and tensile stress in a few seconds and relaxed almost to 0 within two
strength (Fig. 1f and Table S2†). From the thermogravimetric minutes (Fig. 2a); generally, the τ* increased with the increase
analysis (TGA), all the PMMCs presented initial degradation of the cross-linker CDVE content. Take PMMC-5 as an
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

temperature Td5% (5% weight loss) above 217 °C (Fig. 1g and example, the relaxation rate increased with the relaxation
Table S1†). The PMMCs showed a two-stage decomposition temperature (Fig. 2b). The stress relaxation for the PMMCs
trend. We suspect that the weight loss in the first stage was occurred by virtue of the dynamic hemiacetal ester exchange;
caused by the dissociation of the hemiacetal ester bond, and therefore, the Ea from the stress relaxation would be the
the weight loss in the second stage corresponded to the characteristic of the hemiacetal ester exchange process in the
decomposition of the main material. Finally, PMMCs were PMMCs (Fig. 2c). As expected, the Eas of the hemiacetal ester
completely decomposed. DCPNs were lower than those of the acetal DCPNs,40–43 which

Fig. 2 Malleability and reprocessing of the PMMCs. (a) Stress relaxation of the PMMCs at 190 °C. (b) Stress relaxation of the PMMC-5 at different
temperatures. (c) Arrhenius analysis of the characteristic relaxation time τ* versus T for the PMMCs and their activation energies. (d) Appearance of
the PMMC-5 before and after compression remolding at 160 °C under 10 MPa pressure for different times. (e) Extrusion of the PMMC-5 on a plas-
tometer at 200 °C. (f ) Extrusion of the PMMC-5 on a twin-screw mini-extruder at 200 °C at a rotation speed of 15 rpm. (g) FTIR spectra and (h)
representative tensile stress–strain curves of the original and reprocessed PMMC-5. (i) Frequency sweep measurements at different temperatures of
the PMMC-10.

Green Chem. This journal is © The Royal Society of Chemistry 2021


View Article Online

Green Chemistry Paper

means that hemiacetal ester linkage required lower energy to screw mini-extruder, the processing conditions of the hemiace-
activate the exchange reaction than acetal linkage. As a result tal ester DCPNs were first explored with an extrusion plas-
of the rapid relaxation at elevated temperatures, the hemiacetal tometer. PMMC-5 as a typical example had an outflow mass of
ester DCPNs possessed excellent and diverse reprocessability. 5.28 g within 10 min under a pressure of 3.8 kg at 200 °C, and
For the sake of determining the compression remolding after hot pressing, the extruder became a colorless and trans-
efficiency of the hemiacetal ester DCPNs, the samples were hot parent film again (Fig. 2e and Fig. S16†). Then, the continuous
pressed at 10 MPa and 160 °C for different times (Fig. 2d and reprocessing of the hemiacetal ester DCPNs was carried out
Fig. S14†). The pieces of the PMMCs could be recovered into with a twin-screw mini-extruder at 200 °C (Fig. 2f ). During
complete films within 10 min, while the cross-link density was extrusion, viscous flow was observed for the PMMC-5 at
slightly decreased which could be reflected by the decreased 200 °C. No further flow was observed when the applied heat
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

modulus after compression remolding (Fig. S15†). In addition and pressure were removed, which enabled a good retention of
to hot-press recovery, the hemiacetal ester DCPNs show tre- the obtained shape. As can be seen from the FTIR spectra
mendous potential in continuous-extrusion reprocessing. The (Fig. 2g) and tensile properties (Fig. 2h, Fig. S17–S20 and
melt flow index is used to characterize the viscous flow charac- Table S3†) of the original and reprocessed samples, the chemi-
teristics of the polymer in the molten state. To guide the pro- cal structure was maintained well and there is slight reduction
cessing conditions of the hemiacetal ester DCPNs in a twin- in tensile strength and elongation after the reprocessing re-

Fig. 3 Hemiacetal ester exchange reactions. (a and d) Metathesis reactions between the model hemiacetal ester 1 and 2 (a) without PPA (acid) and
(d) with PPA. (b and e) Normalized gas chromatograms of the model hemiacetal ester 1 and 2 mixture before and after metathesis (b) without PPA
and (e) with PPA at different temperatures. (c and f ) Metathesis kinetics of the model hemiacetal ester 1 and 2 (c) without PPA at 120 °C and (f ) with
PPA at 60 °C. (g) Gas chromatograms of 1, 2, 3 and 4. (h) Dissociation of the model hemiacetal ester 1 at 120 °C for 3 h, monitored by 1H NMR. (i)
Fraction of 1 as a function of time during the exchange experiment at different temperatures. ( j) Plot of −ln ([1]/[1]0) versus t for 1, kexp at each temp-
erature was determined by its slope value. (k) Arrhenius analysis of the rate constant kexp versus T for the exchange reaction of 1 and 2.

This journal is © The Royal Society of Chemistry 2021 Green Chem.


View Article Online

Paper Green Chemistry

cycling. In the frequency sweep measurements (Fig. 2i), there the model exchange reaction. They were synthesized by the
is a significant decrease in storage modulus with increasing addition reaction of carboxylic acid and vinyl ether, and their
temperature for PMMC-10, and the storage modulus of chemical structures were characterized in detail by 1H NMR
PMMC-10 also decreased during the time sweep measurement and 13C NMR (Fig. S2–S9†). GC-MS was used to monitor the
at elevated temperatures (for example, at 120 °C) (Fig. S21†). exchange reactions. The small-molecule model compounds
These demonstrate the dissociative exchange mechanism of were well separated on a gas chromatogram (GC) (Fig. 3g) and
the hemiacetal ester linkage. This rapid dissociation and readily identified by mass spectrometry (MS) because each
rebinding of the hemiacetal ester structure contributed to the component possesses a distinct molecular weight (Fig. S22†).
rapid continuous reprocessing of the networks. The exchange reactions of small-molecule model hemiacetal
ester compounds without carboxylic acid (Fig. 3a) and with
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

3.4. Dynamic exchange mechanism of the hemiacetal ester carboxylic acid (Fig. 3d) were both investigated. Fig. 3b shows
DCPNs via small-molecule model reactions and theoretical the gas chromatograms of model systems without carboxylic
calculation acid before and after reactions, and it can be found that the
To demonstrate the dynamic exchange mechanism of the exchange reaction could reach equilibrium after 3 h at 120 °C
hemiacetal ester DCPNs, we utilized small-molecule model in an acid-free system, while it took shorter time to reach equi-
compounds IBA–CVE (1) and PPA–IVE (2) to explore the librium at 120 °C and could occur at lower temperatures in a
exchange reaction, and we also utilized IBA–IVE (3) and PPA– carboxylic acid-containing system (Fig. 3e). A quantitative
CVE (4) to confirm that they are indeed the products formed in kinetic study was carried out by considering the total area frac-

Fig. 4 Theoretical calculation of the hemiacetal ester exchange reaction. (a) Dynamic exchange mechanism of the hemiacetal ester. (b) DFT-calcu-
lated energy surface for the exchange reaction between IBA–CVE and PPA–IVE.

Green Chem. This journal is © The Royal Society of Chemistry 2021


View Article Online

Green Chemistry Paper

tion of the model hemiacetal ester compounds (1) and (3) as a DCPNs combining fast reprocessability and high performance,
function of time. In the case of the acid-free system, the which is beneficial for the industrialization of DCPNs.
exchange was slow at first, then became fast (Fig. 3c and
Fig. S23†). By contrast, the exchange was the fastest at the
beginning for the carboxylic acid-containing system (Fig. 3f
and Fig. S24†). This manifests that carboxyl group could accel- Conflicts of interest
erate the exchange reaction of hemiacetal esters. The kinetics The authors declare no competing financial interest.
of the exchange reaction between IBA–CVE and PPA–IVE at
different temperatures were also investigated by monitoring
the fraction of PPA–IVE in the reaction system (Fig. 3i). The
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

fitting function showed that the exchange reaction was in Acknowledgements


accordance with a pseudo-second-order kinetic reaction
(Fig. 3j). Accordingly, the relationship between the reaction We thank the financial support from the National Natural
rate constant k and the temperature obeyed the Arrhenius Science Foundation of China (no. 52073296 and 51773216),
equation, and the activation energy of the reaction was calcu- Research Project of Technology Application for Public Welfare
lated to be about 56 kJ mol−1 (Fig. 3k), and the details are of Ningbo City (no. 202002N3091), and Youth Innovation
described in Fig. S25–S27.† To further certify the exchange Promotion Association, CAS (no. 2018335).
mechanism of hemiacetal ester, the model hemiacetal ester
compound (1) was heated at 120 °C for 3 h, and the reaction
was monitored using 1H NMR (Fig. 3h). The appearance of Notes and references
signals for carboxyl and cyclohexyl vinyl ether is indicative of
the dissociative mechanism, which is in agreement with the 1 Z. P. Zhang, M. Z. Rong and M. Q. Zhang, Prog. Polym. Sci.,
result from the frequency sweep measurements (Fig. 2i). 2018, 80, 39–93.
In addition, we had studied the exchange process in more 2 S. Ma and D. C. Webster, Prog. Polym. Sci., 2018, 76, 65–
depth through theoretical calculations. Fig. 4a clearly shows 110.
that the dissociative exchange pathway involves cleavage of the 3 H. Sardon and A. P. Dove, Science, 2018, 360, 380–381.
hemiacetal ester linkages into intermediate vinyl ether groups 4 P. R. Christensen, A. M. Scheuermann, K. E. Loeffler and
and carboxyl groups, then re-connected via the addition reac- B. A. Helms, Nat. Chem., 2019, 11, 442–448.
tion to produce new hemiacetal ester linkages. Fig. 4b and 5 J. M. Garcia, G. O. Jones, K. Virwani, B. D. McCloskey,
Table S4† exhibit the Gibbs free energy and transition state D. J. Boday, G. M. ter Huurne, H. W. Horn, D. J. Coady,
geometry of hemiacetal ester exchange with the assistance of A. M. Bintaleb, A. M. Alabdulrahman, F. Alsewailem,
DFT calculations (B3LYP-D3BJ/6-31G(d,p) level). For the system H. A. Almegren and J. L. Hedrick, Science, 2014, 344, 732–
with carboxylic acid, its Gibbs free energy is lower than the 735.
acid-free system, which proves that the proton from the car- 6 W. Post, A. Susa, R. Blaauw, K. Molenveld and
boxylic acid could catalyze the dissociation of the hemiacetal R. J. I. Knoop, Polym. Rev., 2019, 60, 359–388.
ester and further accelerate the exchange rate. 7 B. Wang, S. Ma, S. Yan and J. Zhu, Green Chem., 2019, 21,
5781–5796.
8 C. J. Kloxin, T. F. Scott, B. J. Adzima and C. N. Bowman,
4. Conclusions Macromolecules, 2010, 43, 2643–2653.
9 Y. Jin, Z. Lei, P. Taynton, S. Huang and W. Zhang, Matter,
In summary, the dynamic feature of hemiacetal ester was suc- 2019, 1, 1456–1493.
cessfully exploited to produce DCPNs with fast reprocessability 10 M. Capelot, M. M. Unterlass, F. Tournilhac and L. Leibler,
and high performance. A facile and green method combining ACS Macro Lett., 2012, 1, 789–792.
free radical copolymerization and dynamic cross-linking reac- 11 M. Röttger, T. Domenech, R. van der Weegen, A. Breuillac,
tion—in situ polymerization and dynamic cross-linking R. Nicolaÿ and L. Leibler, Science, 2017, 356, 62–65.
(ISPDC)—was explored to synthesize the hemiacetal ester 12 W. Zou, J. Dong, Y. Luo, Q. Zhao and T. Xie, Adv. Mater.,
DCPNs. The hemiacetal ester exchange proceeded extremely 2017, 29, 1606100.
fast and followed the dissociative mechanism at elevated temp- 13 N. Zheng, Y. Xu, Q. Zhao and T. Xie, Chem. Rev., 2021, 121,
eratures. Moreover, the excess carboxyl group in the hemiacetal 1716–1745.
ester DCPNs could significantly accelerate the exchange rate of 14 D. Montarnal, M. Capelot, F. Tournilhac and L. Leibler,
hemiacetal esters, corresponding to their ultra-fast stress relax- Science, 2011, 334, 965–968.
ation and extremely short reprocessing time at elevated temp- 15 X. X. Chen, M. A. Dam, K. Ono, A. Mal, H. B. Shen,
eratures. As a result, the hemiacetal ester DCPNs could be S. R. Nutt, K. Sheran and F. Wudl, Science, 2002, 295, 1698–
reprocessed by extrusion on a twin-screw extruder (continuous) 1702.
as well as subjected to compression remolding (incontinuous). 16 S. Yu, R. Zhang, Q. Wu, T. Chen and P. Sun, Adv. Mater.,
This work disclosed a facile and green method to produce 2013, 25, 4912–4917.

This journal is © The Royal Society of Chemistry 2021 Green Chem.


View Article Online

Paper Green Chemistry

17 A. Ruiz de Luzuriaga, R. Martin, N. Markaide, A. Rekondo, 36 N. Van Herck, D. Maes, K. Unal, M. Guerre, J. M. Winne
G. Cabanero, J. Rodriguez and I. Odriozola, Mater. Horiz., and F. E. Du Prez, Angew. Chem., 2020, 59, 3609–3617.
2016, 3, 241–247. 37 H. Ying, Y. Zhang and J. Cheng, Nat. Commun., 2014, 5,
18 L. Imbernon, E. K. Oikonomou, S. Norvez and L. Leibler, 3218.
Polym. Chem., 2015, 6, 4271–4278. 38 W.-X. Liu, C. Zhang, H. Zhang, N. Zhao, Z.-X. Yu and J. Xu,
19 C. D. Roy and H. C. Brown, Monatsh. Chem., 2007, 138, J. Am. Chem. Soc., 2017, 139, 8678–8684.
879–887. 39 D. Sang, J. Wang, Y. Zheng, J. He, C. Yuan, Q. An and
20 J. J. Cash, T. Kubo, A. P. Bapat and B. S. Sumerlin, J. Tian, Synthesis, 2017, 49, 2721–2726.
Macromolecules, 2015, 48, 2098–2106. 40 Q. Li, S. Ma, S. Wang, W. Yuan, X. Xu, B. Wang, K. Huang
21 O. R. Cromwell, J. Chung and Z. Guan, J. Am. Chem. Soc., and J. Zhu, J. Mater. Chem. A, 2019, 7, 18039–18049.
Published on 08 October 2021. Downloaded by University of Leeds on 11/14/2021 9:51:06 PM.

2015, 137, 6492–6495. 41 Q. Li, S. Ma, S. Wang, Y. Liu, M. A. Taher, B. Wang,


22 L. Shen, J. Cheng and J. Zhang, Eur. Polym. J., 2020, 137, K. Huang, X. Xu, Y. Han and J. Zhu, Macromolecules, 2020,
109927. 53, 1474–1485.
23 Y. X. Lu, F. Tournilhac, L. Leibler and Z. Guan, J. Am. Chem. 42 Q. Li, S. Ma, P. Li, B. Wang, H. Feng, N. Lu, S. Wang, Y. Liu,
Soc., 2012, 134, 8424–8427. X. Xu and J. Zhu, Macromolecules, 2021, 54, 1742–1753.
24 P. Taynton, K. Yu, R. K. Shoemaker, Y. Jin, H. J. Qi and 43 W. L. Peng, Y. You, P. Xie, M. Z. Rong and M. Q. Zhang,
W. Zhang, Adv. Mater., 2014, 26, 3938–3942. Macromolecules, 2020, 53, 584–593.
25 M. Harada, J. Ando, M. Yamaki and M. Ochi, J. Appl. Polym. 44 M. Delahaye, F. Tanini, J. O. Holloway, J. M. Winne and
Sci., 2015, 132, 41296. F. E. Du Prez, Polym. Chem., 2020, 11, 5207–5215.
26 M. G. Mohamed, E. C. Atayde, B. M. Matsagar, J. Na, 45 M. Rottger, T. Domenech, R. van der Weegen,
Y. Yamauchi, K. C. W. Wu and S.-W. Kuo, J. Taiwan Inst. A. B. R. Nicolay and L. Leibler, Science, 2017, 356, 62–65.
Chem. Eng., 2020, 112, 180–192. 46 J. Qiu, S. Ma, S. Wang, Z. Tang, Q. Li, A. Tian, X. Xu,
27 M. Guerre, C. Taplan, R. Nicolay, J. M. Winne and F. E. Du B. Wang, N. Lu and J. Zhu, Macromolecules, 2021, 54, 703–
Prez, J. Am. Chem. Soc., 2018, 140, 13272–13284. 712.
28 C. Taplan, M. Guerre, J. M. Winne and F. E. Du Prez, Mater. 47 S. Wang, S. Ma, J. Qiu, A. Tian, Q. Li, X. Xu, B. Wang,
Horiz., 2020, 7, 104–110. N. Lu, Y. Liu and J. Zhu, Green Chem., 2021, 23, 2931–2937.
29 W. Denissen, I. De Baere, W. Van Paepegem, L. Leibler, 48 D. Boucher, J. Madsen, L. Yu, Q. Huang, N. Caussé,
J. Winne and F. E. Du Prez, Macromolecules, 2018, 51, 2054– N. Pébère, V. Ladmiral and C. Negrell, Macromolecules,
2064. 2021, 54, 6772–6779.
30 W. Denissen, M. Droesbeke, R. Nicolaÿ, L. Leibler, 49 P. J. Stephens, F. J. Devlin, C. F. Chabalowski and
J. M. Winne and F. E. Du Prez, Nat. Commun., 2017, 8, 14857. M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627.
31 M. Pepels, I. Filot, B. Klumperman and H. Goossens, 50 S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem.
Polym. Chem., 2013, 4, 4955–4965. Phys., 2010, 132, 154104.
32 M. M. Obadia, B. P. Mudraboyina, A. Serghei, D. Montarnal 51 P. C. Hariharan and J. A. Pople, Theor. Chim. Acta, 1973, 28,
and E. Drockenmuller, J. Am. Chem. Soc., 2015, 137, 6078– 213–222.
6083. 52 W. J. Hehre, R. Ditchfield and J. A. Pople, J. Chem. Phys.,
33 C. A. Tretbar, J. A. Neal and Z. Guan, J. Am. Chem. Soc., 1972, 56, 2257–2261.
2019, 141, 16595–16599. 53 A. V. Marenich, C. J. Cramer and D. G. Truhlar, J. Phys.
34 Y. Nishimura, J. Chung, H. Muradyan and Z. Guan, J. Am. Chem. B, 2009, 113, 6378–6396.
Chem. Soc., 2017, 139, 14881–14884. 54 S. Ma and D. C. Webster, Macromolecules, 2015, 48, 7127–
35 H. Lei, S. Wang, D. J. Liaw, Y. Cheng, X. Yang, J. Tan, 7137.
X. Chen, J. Gu and Y. Zhang, ACS Macro Lett., 2019, 8, 582– 55 S. Ma, D. C. Webster and F. Jabeen, Macromolecules, 2016,
587. 49, 3780–3788.

Green Chem. This journal is © The Royal Society of Chemistry 2021

You might also like