Vanillin-Based Polyschiff Vitrimers ... - ACS Publications

Sep 28, 2018 - epoxy resin, one of the most important thermosets, has also been prepared by epoxidation .... Canon EOS 1100D, using 2× and 5× object...
0 downloads 0 Views 692KB Size
Subscriber access provided by University of Sunderland

Article

Vanillin-based polyschiff vitrimers: reprocessability and chemical recyclability Hongwei Geng, Yuli Wang, Qingqing Yu, Shaojin Gu, Yingshan Zhou, Weilin Xu, Xi Zhang, and De-zhan Ye ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b03925 • Publication Date (Web): 28 Sep 2018 Downloaded from http://pubs.acs.org on September 28, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

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

ACS Sustainable Chemistry & Engineering

1

Vanillin-based polyschiff vitrimers: reprocessability and chemical

2

recyclability

3

Hongwei Geng c, Yuli Wang c, Qingqing Yu c, Shaojin Gu c, Yingshan Zhou c, Weilin Xu a, Xi Zhang b, Dezhan Ye a,b,c*

4

a

State Key Laboratory Cultivation Base for New Textile Materials & Advanced Processing Technology , Wuhan

5

6

Textile University, No. 1 Yangguang Avenue, Jiangxia District, Wuhan, Hubei 430200, China.

b

State Key Laboratory of Polymer Material Engineering, Sichuan University, No. 24 South Section 1, Yihuan

7

8

Road, Chengdu 610065, China.

c

School of Materials Science and Engineering, No. 1 Yangguang Avenue, Jiangxia District, Wuhan, Hubei

9

430200, China.

10

*Corresponding authors: [email protected]

11

ABSTRACT

12

In this study, dynamic imine covalent bonds were introduced into vanillin-based vitrimers networks,

13

endowing thermosets with hot-reprocessing ability and chemical recyclability under acid hydrolysis. First,

14

dialdehyde monomer, which was synthesized from lignin-derived vanillin monomer, was reacted with

15

conventional amine cross-linkers to form dynamic imine bond networks. Even after three hot-processing cycles,

16

the tensile strength and elongation at break of polyschiff vitrimers could be recovered at least up to 71.2% and

17

72.8%, respectively, through the imine metathesis reaction. Importantly, the dialdehyde monomers showed

18

enhanced recyclability under strong acid solution, and could be reused to regenerate polyschiff vitrimers. These

19

characteristics of reprocessablility, recyclablility, and bio-based monomer present a feasible way to satisfy the

20

demands of sustainability.

21

Keywords: vanillin, bio-based thermoset, dynamic covalent bond, imine, self-healing, reprocessing, chemical

22

recycling 1

ACS Paragon Plus Environment

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

23

Page 2 of 21

Introduction

24

To address sustainability challenges caused by petrochemical depletion, researchers are exploiting

25

renewable resources to generate bio-based chemicals, materials, and biofuels. Lignin, a byproduct of pulp and

26

paper refining, shows great potential to be a desirable alternative to petroleum feedstock if we consider its

27

abundance as a natural resource 1. To take advantage of its renewability and backbone rigidity, lignin-based

28

thermosets, such as epoxy and phenol resin , have been extensively studied.

29

2

3

Two approaches, fractionation and chemical modification (e.g., anhydride reaction and epoxidation), have 4

30

been traditionally adopted for preparing lignin-derived thermosets

31

deconstruction of its aromatic structures is challenging because of its poor reactivity and solubility. Compared

32

with these methods, the preparation of bio-based thermosets from well-characterized lignin-derived monomers

33

is desirable and has received extensive attention. Vanillin, a renewable and aromatic compound from lignin

34

and its derivatives (e.g., eugenol, 4-Hydroxy-3-methoxybenzyl alcohol, and vanillic acid) have been widely

35

reported for producing epoxy and phenolic resin

36

reaction to synthesize UV-curable monomers for coating applications

37

most important thermosets, has also been prepared by epoxidation of vanillyl alcohol

38

vanillin, and vanillic acid derivations

39

monomer in the synthesis of high-performance of benzoxazines resin

40

obtained by vanillin .

13

8-11

. However, using lignin without

5-7

,

. For example, eugenol has been modified by thiolene 12

. Interestingly, epoxy resin, one of the 10

, dihydroeugenol 9,

. Recently, vanillin and guaiacol have been utilized as bio-based aromatic 14

. Moreover, phenolic resin also can be

3

41

Traditionally, thermosets have permanent covalent bond networks that preclude the reprocessing of cured

42

polymers. Designing cross-linked networks that can be reprocessed, self-healed, and reshaped is still a challenge

43

to enable available manufacturing techniques for thermoset materials. One approach to solving these limitations

44

is to introduce the dynamic covalent into thermoset structures . For example, Leibler et al. prepared a malleable

15

2

ACS Paragon Plus Environment

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

ACS Sustainable Chemistry & Engineering

16

45

polyester thermoset based on transesterification reactions at high temperatures with metal catalysts

. Jianjun

46

Cheng et al. designed hindered urea bonds and used them to make polyureas and poly(urethane-urea)s

47

cross-linked networks that can self-heal at room temperature . Other reversible reactions, such as retro-Diels–

48

Alder , disulfide metathesis , boronic ester exchange , oxime−carbamate , and transcarbamoylation

49

also been adopted in the preparation of reprocessable and self-healing cross-linked polymeric materials.

50

Considering the superior renewability of lignin in natural plants, it is desirable to prepare lignin-based thermosets

51

with intrinsically reprocessable, recyclable, and self-healable ability if one is concerned about the impact of the

52

depletion of petrochemical resources. However, few studies have been performed in this emerging field so far.

53

Very recently, Jinwen Zhang reported a unique eugenol-derived epoxy that was cured with succinic anhydride .

54

Under the catalysis of zinc acetylacetonate hydrate, this cross-linked polymer exhibited excellent shape-changing

55

and crack-healing ability. They also prepared lignin-based vitrimers from ozone-treated Kraft lignin and sebacic

56

acid epoxy, showing excellent repair properties at elevated temperatures .

17

18

19

20

21

22

have

23

24

57

The imine bond is a classic dynamic covalent bond, which is formed by the condensation reaction between

58

aldehydes and amines. Three equilibrium processes are involved in the imine reactions, including imine

59

condensation/hydrolysis, imine exchange, and imine metathesis

60

widespread use as self-healing hydrogel

61

resins. Recently, imine chemistry has been designed to prepare ultra-thin solid-state Li-Ion electrolyte membranes

62

and repairable woven carbon fiber composites, in which the dialdehyde is terephthaldehyde

63

exchange reactions of imine, the reported repairable polyimine could be dissolved/recycled by using amine

64

monomer. Benezene-1,3, 5-tricarbaldehyde also has been used as a functional cross-linker to form dynamic imine

65

bonds in self-healing and reprocessable poly(dimethyl siloxane) elastomer

66

regard to vanillin-derived Schiff base materials, their utilizations are usually limited to the fields of water

26-28

25

and drug-release carriers

. Imine chemistry, traditionally finding 29,30

3

ACS Paragon Plus Environment

has been under-used for thermoset

33

31,32

. Owing to the

and polybutadiene rubber

34

. With

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

67

treatment because of their excellent chelation with metal ions

35,36

.

68

In this study, cross-linked polyimine/polyschiff vitrimers from a lignin-based monomer, vanillin, were

69

prepared by employing imine chemistry. These thermosets could be reprocessed and self-repaired under elevated

70

temperatures, because of the thermal reversibility of imine bonds. Interestingly, the chemical recycling of

71

dialdehyde monomer was demonstrated under acid solution at moderate temperature (50 °C). Moreover, the

72

recycled dialdehyde monomer can be reused in the synthesis of polyschiff vitrimers. This work may provide some

73

new insights in the design of recyclable lignin-based vitrimers with reprocessable and self-healing properties.

74

Experimental

75

Materials

76

Vanillin (≥ 99%), tris(2-aminoethyl)amine (≥ 96%), diethylenetriamine (≥ 99%) and 1,4-dibromo butane (≥

77

98%) were purchased from Aladdin company. All other analytical reagents were commercial chemicals and used

78

without further purification.

79

Synthesis of dialdehyde monomer from vanillin (DAV)

80

Vanillin (12 mmol, 2.4 eq), anhydrous K2CO3 (12 mmol, 2.4 eq), and CH3CN (30 mL) were added to a 150 mL

81

three-necked bottle under magnetic stirring. Then, the mixture was degassed with nitrogen for at least 30 min,

82

and the reaction was started by adding 5 mmol (1.0 eq) 1,4-dibromo butane. The reaction was conducted for 24 h

83

under reflux. Finally, the reaction mixture was poured into a large volume of water to remove salt. After filtration,

84

the precipitation was dissolved into CH2Cl2 and was washed three times with NaOH (1 M) and distilled water. The

85

organic phase was dried over anhydrous MgSO4 before the solvent was condensed by a rotary evaporator in

86

vacuum. After being vacuum dried in an oven overnight, DAV with a yield of 86% was obtained as a white powder.

87

1

H-NMR (400 MHz, CDCl3) δ (ppm): 9.76 (s, 2H, Ar-CHO), 7.36 (d, 2H, Ar-H), 7.34 (s, 2H, Ar-H), 6.90 (d, 2H, 4

ACS Paragon Plus Environment

Page 4 of 21

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

ACS Sustainable Chemistry & Engineering

13

88

Ar-H), 4.13 (t, 4H, O-CH2-), 3.81 (s, 6H, OCH3), 2.10-1.99 (m, 4H, O-CH2-CH2-). C-NMR (400 MHz, CDCl3) δ (ppm):

89

190.88, 153.92, 149.77, 130.00, 126.69, 111.37, 109.20, 68.64, 55.94, and 25.75.

90

Polyimine cross-linked film preparation

91

The cross-linked films were prepared by dissolving DAV in an appropriate amount of CH2Cl2 solvent; then, a

92

precise amount of diethylenetriamine and tris(2-aminoethyl)amine were added into the solution under stirring

93

conditions at 50 °C for 10 min. Immediately, the mixture was poured into a square PTFE tray, and a glass plate was

94

used to cover it, which was then left at room temperature overnight. The obtained defect-free film was further

95

cured in a drying oven at 50 °C for at least 24 h. Aimed at obtaining polyschiff vitrimers with different crosslinking

96

degrees, the molar ratio of tris(2-aminoethyl)amine crosslinker to diethylenetriamine monomer was adjusted.

97

The details for preparing the cross-linked polyimine films are recorded in Table 1.

98

Table 1. Predetermined content of DAV and amine monomers used for preparing cross-linked polyschiff films

DAV

diethylenetriamine

tris(2-aminoethyl)amine

Sample 1

6mmol, 2.1504 g

1.8 mmol, 0.1974 ml

2.8 mmol, 0.437 ml

Sample 2

6 mmol, 2.1504 g

3.0 mmol, 0.329 ml

2 mmol, 0.3122 ml

Sample 3

6 mmol, 2.1504 g

--

4 mmol, 0.6244 ml

99

100

Characterization

101

Fourier transform infrared (FTIR) spectra were recorded on a Nicolet 560 FTIR spectrometer. The powder

102

samples were characterized using the KBr pellet technique. With respect to the polyschiff films, they were directly

103

measured by the ATR model. Each spectrum was recorded with 32 scans in the frequency range of 4000–400 cm

104

with a resolution of 2 cm−1.

5

ACS Paragon Plus Environment

−1

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

105

106

Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker 400 (400 MHz) spectrometer, using

CDCl3 as a solvent.

107

The thermal properties of polyimine film were performed on a differential scanning calorimeter (DSC)

108

(NETZSCH, DSC 214). The sample was dried in a vacuum oven at 100 °C for 3 h before conducting experiments.

109

First, the sample was heated from room temperature to 105 °C and kept at equilibrium in this state for 3 min to

110

remove the water. Then, the temperature was immediately dropped to −40 °C before being gradually increased to

111

200 °C at 10 °C/min. All the abovementioned temperature controls were performed under nitrogen atmosphere

112

(60 mL/min flow rate).

113

The thermal stability of the films was measured by thermal gravimetric analysis (TGA; TA instrument, model

114

Q-500) at a heating rate of 10 °C/min from room temperature to 600 °C under nitrogen atmosphere (50 mL/min

115

flow rate).

116

117

Dynamic mechanical analysis (DMA) was performed on a Q 800 instrument using single-beam mode at 1 Hz

with a heating rate of 3 °C/min in air atmosphere.

118

The mechanical properties of the polyschiff cross-linked films were measured on a 5567 Universal Testing

119

machine with a 1 kN sensor. The gauge length was 50 mm, and the test was performed at a 20 mm/min

120

cross-head speed. The average thickness values of the films with 1 cm width were calculated with a micrometer

121

over five measurements. At least eight species were tested for each sample.

122

123

Optical microscopy images, which were used to monitor crack healing, were obtained with an Olympus BH-2

equipped with a Canon EOS 1100D, using 2× and 5× objectives.

124

The self-healing ability of dumbbell-shaped samples was tested by cutting the center of each sample using

125

scissors, after which the cut pieces were gently pressed by a 200 g weight at 150 °C for 1 h. After cooling, the

126

pieces were subjected to strain-stress experiments to test the recovery of mechanical properties.

6

ACS Paragon Plus Environment

Page 6 of 21

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

ACS Sustainable Chemistry & Engineering

127

A reprocessing test was performed on a hot-pressing machine. The films were cut into small pieces with

128

scissors, placed into a round steel mold (10 cm diameter and 0.5 mm thickness), and reprocessed at 150 °C for 10

129

min at 10 MPa. After cooling to room temperature, the recycled films were obtained and stress−strain tests were

130

performed as previously described.

131

Chemical recycling of cross-linked films was performed by immersing the film in HCl solution (pH = 2, 4, 6);

132

then, the insoluble residuals were collected, vacuum-dried, and characterized by FTIR and 1H-NMR methods. The

133

weighted residuals were used as recycled DAV monomer to prepare cross-linked polyimine films as previously

134

described.

135

Results and discussion

136

Synthesis and characterization of polyschiff films

137

The preparation of cross-linked polyimine vitrimers is schematically illustrated in Figure 1. First, dialdehyde

138

was synthesized from vanillin and 1,4-dibromo butane; then, it was condensed with amine monomers to obtain a

139

light yellow film.

140 141

Figure 1. Synthetic route of polyschiff thermosets from vanillin

142

143

Dialdehyde monomer was successfully prepared as shown in the FTIR spectra (Supporting Information Figure

144

S1). The absorption peak around 3165.1 cm , corresponding to the vibration of the vanillin–OH group,

−1

7

ACS Paragon Plus Environment

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

145

disappeared after the reaction with 1,4-dibromo butane. The new vibration of methyl and methylene groups near

146

2974.7, 2944.7, and 2858.5 cm also demonstrated the successful synthesis of DAV. Furthermore, the absorption

147

peak of aromatic aldehyde, centered near 1670 cm , was also observed. As shown in the NMR spectra of Figure

148

S2, the absorption signal for the methylene proton, which is adjacent to the phenolic ether (Ph-O-CH2-) group,

149

was detected at 4.13 and 68.64 ppm, indicating that the dialdehyde monomer was successfully synthesized from

150

vanillin. Moreover, the absorption peak at around 9.76 and 190.88 ppm is confirmed as the aldehyde group.

−1

−1

151

Because of the fast imine condensation reaction, cross-linked polyimine films were easily prepared through

152

the reaction between the dialdehyde (DAV), diethylenetriamine, and tris(2-aminoethyl)amine) monomer even at

153

room temperature. The mixed solution became light yellow from colorless in only 10 min at 50 °C, suggesting the

154

beginning of the imine reaction. FTIR results were used to prove the formation of Schiff structures in the film

155

(Supporting Information Figure S3). A new emerging peak near 1640 cm and a disappearing absorption at 1678

156

cm , correlating with imine and aromatic aldehyde, respectively, strongly suggests the formation of imine

157

bonding networks . A gel fraction experiment was conducted to check the imine networks. After being refluxed

158

with CH2Cl2 at 50 °C for 12 h, the bulk film of sample 3 still retained its original shape without breaking into small

159

particles, with approximately 96% residual weight percentage.

−1

−1

32

160

The thermal properties of polyschiff films were studied here in order to confirm the thermal reprocessing

161

and self-healing conditions. The results shown in Figure S4 indicate that the glass transition temperature (Tg) of

162

polyimine films increased with the cross-linking degree, which rose from 48 °C to 64 °C. Furthermore, their

163

thermal stabilities were characterized by the TGA method as illustrated in Figure S5. Clearly, all films are thermally

164

stable up to a temperature of at least 300 °C (onset decomposition temperature, To). The maximum

165

decomposition temperature (Tp) ranged from 349 to 368 °C, which is obviously associated with the cross-linking

166

degree. Finally, the residual weight percentages of polyimine films are still almost 36% up to 600 °C. All these

8

ACS Paragon Plus Environment

Page 8 of 21

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

ACS Sustainable Chemistry & Engineering

167

results demonstrate that polyschiff networks are thermally stable under conventional modeling approaches. The

168

detailed results of thermal properties of polyschiff vitrimers are recorded in Table 2.

169

170

Table 2. Summary of the thermal properties of polyschiff films

a

Toa (oC)

Tpb (oC)

Residual weight (wt %) at 600 oC

Tg (oC) DSC

νec (mol/m 3)

Pristine Sample 1

315

355

36.1

60

311

Pristine Sample 2

301

349

37.4

48

--

Pristine Sample 3

335

368

35.2

64

1120

Initial decomposition temperature; b maximum decomposition temperature; c calculated from rubbery modulus measured by DMA experiments

171

172

Figure 2 illustrates the dynamic mechanical properties of polyschiff thermosets. The storage modulus of

173

pristine 1 (E′ = 1713 MPa) is lower than that of pristine 3 (E′ = 2341 MPa) at 25 °C, which is obviously attributed to

174

the higher crosslinking degree in sample 3. The tanδ peak is often used to determine the Tg. It was noticed that

175

the Tg of crude samples 3 and 1 are 76 and 75 °C, respectively, as shown in Figure 2b. A rubbery plateau can be

176

observed above the Tg temperature for both, indicating the presence of cross-linking networks in this system. The

177

cross-linking density (νe) of virginal samples 3 and 1 are 1120 and 311 mol/m , respectively, according to the

178

previously reported method 37.

3

179 180

Figure 2. (a) Storage Modulus and (b) tanδ of polyschiff thermosets from vanillin

181

182

The mechanical properties of pristine polyimine vitrimers were tested as illustrated in Figure S6. Overall, the

183

tensile strength and elongation at break of polyschiff vitrimers ranged from 47.43 ± 2.46 MPa to 57.1 ± 6.01 MPa, 9

ACS Paragon Plus Environment

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

Page 10 of 21

184

and 16.34 ± 2.86% to 13.01 ± 1.51%, respectively, obviously showing no large differences. The polyschiff vitrimers

185

in this work are much tougher than those of recently reported polyimine vitrimers, whose stress and elongation

186

at break are approximately 40 MPa and 5%, respectively. In contrast with the reported rigidity of

187

terephthalaldehyde, the aromatic aldehyde group is adjacent to the butyl group, resulting in the flexible structure

188

of DAV 31,32.

189

Dynamic properties

190

Self-healing ability

191

Traditional thermosets do not reform their bulk integrity once they are damaged or scratched; this problem 38

39,40

192

can be resolved by introducing dynamic bonds into networks, such as Diels-alder adduct , disulfide linkage

193

and dioxaborolane

194

condensation, exchange, and metathesis

195

thermosets. With the expected behaviors of dynamic imine bonds, the self-healing properties of polyschiff

196

vitrimers were first evaluated by a scratch recovery test. The scratched film was subjected to a heating stage at

197

180 °C and monitored using optical microscopy, as shown in Figure 3. Obviously, the scratch immediately became

198

shallower in only 15 min, and it finally disappeared after 2 h, indicating its self-healing ability at the evaluated

199

temperature.

200

41

,

. With regard to the imine bond, three types of reverse imine reactions, including imine 25

, are obviously desirable for designing the dynamic properties of

It is widely acknowledged that covalent adaptable networks are classified into two groups, dissociative and 42

201

associative, depending on their exchange mechanism

202

excluded, mainly because of the trace amount of water under the high self-healing temperature. This explanation

203

could be proved by the absence of aromatic aldehyde absorption in the FTIR spectra of the first recycled sample 1,

204

as shown in Figure S7. With the equivalent molar ratio of aldehyde and amine in the vitrimers preparation, the

205

self-healing mechanism polyschiff vitrimers does not belong to the imine exchange reaction (residual amine 10

. In this work, the breakage of imine bonds could be

ACS Paragon Plus Environment

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

ACS Sustainable Chemistry & Engineering

34

206

group exchanged with imine)

207

metathesis, as shown in Figure 4.

. Finally, we deduced that its dynamic properties are controlled by imine

208 209

Figure 3. Optical microscopy images of scratched pristine sample 1 before and after healing at 180 °C for different times

210

211 212

Figure 4. Self-healing ability of crosslinking polyschiff vitrimers endowed by dynamic imine metathesis reaction

213

214

To visualize the self-healing properties of polyimine vitrimers, the dog-bone shape of sample 1 was cut with

215

scissors, and the cut pieces were realigned with an overlap of 2 mm × 3 mm with a 250 g weight at 150 °C for 1 h.

216

As shown in Figure 5a, the cut edges were melted together. The Young’s modulus, stress at break, and elongation

217

at break of the original sample 1 are 724.38 ± 43.65 MPa, 51.12 ± 3.99 MPa, and 13.01 ± 1.51%, respectively. A

218

healing time of 1 h results in 74.5% and 118.5% recovery of tensile strength and elongation at break, respectively

219

(Figure 5b).

11

ACS Paragon Plus Environment

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

220 221

Figure 5. (a) Photo of cut sample 1 before and after healing for 1 h at 150 °C under 200 g weight in oven; (b) stress−strain curves of the thermally healed

222

sample

223

224

Mechanical properties recovery

225

Once the self-healing ability of polyschiff vitrimers was confirmed, their reprocessability was studied. First,

226

the polyschiff vitrimers were cut into small pieces and then hot-pressed at 150 °C for 10 min to obtain the bulk

227

materials. These procedures were repeated three times to obtain the dog-bone-shaped samples for tensile-strain

228

experiments.

229 230

Figure 6. (a) Reprocessing experiments of polyimine vitrimers in a hot press; (b) bar chart of Young’s modulus, (c) tensile strength at break, and (d) fractured

231

elongation of the original, first, second, and third cycles of vitrimers obtained using hot processing at 150 °C for 10 min

232 12

ACS Paragon Plus Environment

Page 12 of 21

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

ACS Sustainable Chemistry & Engineering

233

The tensile-strain tests (Figure 6) demonstrated that the stress and elongation at break for samples 3 and 1

234

still kept at the same level of original, and sometimes even slightly higher after three recycles. However, these

235

properties of sample 2 become inferior with increasing hot processing cycles, perhaps because of its faster

236

consumption of dynamic bonds. Furthermore, it was found that the tensile strength and elongation at break is

237

proportional to the crosslinking degree under the same hot-processing batch. Meanwhile, the Young’s modulus

238

shows an increasing trend with the declining cross-linking degree. These results demonstrate that the mechanical

239

properties of polyschiff films are better with the higher cross-linking points in this work. Finally, though sample 2

240

shows the lowest tensile strength and elongation at break after three hot-processing cycles, it still exhibits up to

241

71.2 % and 72.8 % recovery.

242

Chemical recycling of DAV monomer

243

In taking advantage of the unstable Schiff structure under acid conditions

43

, it is interesting to test the

244

chemical recycling of DAV monomer and its reuse in the formation of the cross-linked polyschiff film. Here, the

245

samples were immersed in an acid solution (pH =2, 4, 6) at 50 C for 24 h under continuous stirring. Obviously, the

246

bulk materials break into small particles at a pH of 2, leading to the milky mixtures after acid hydrolysis (Figure 7a).

247

In order to quantitatively evaluate the influence of pH on this chemical recycling process, the residual weight

248

percentages of vitrimers were compared with the theoretical content of DAV in vitrimers, as illustrated in Figure

249

7b. The weight percentage of insoluble substances at a pH of 2 is slightly lower than theoretical content of DAV,

250

indicating complete decomposition. However, the opposite trend was observed at higher pH values. Thus, it could

251

be concluded that a stronger acid condition is beneficial to chemical recycling. Furthermore, under the same

252

conditions, the chemical degradation of sample 3 is much slower in contrast with that of the others, which is

253

clearly associated with the cross-linking degree. These results are attributed to the nature of the dynamic imine

254

reaction, in which reversible equilibrium could be accelerated under acid water solution, thus resulting in the 13

°

ACS Paragon Plus Environment

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

255

breakage of imine bonds and the new formation of aldehyde structures.

256

In order to test the film reforming ability, the white powder was recovered by filtration and its structure was

257

characterized by the FTIR and H-NMR methods, as shown in Figure 8. Compared with the original vanillin-based

258

dialdehyde, the H-NMR spectrum of the recycled dialdehyde does not show any differences in the peak positions

259

and area integral of hydrogen protons. The H-NMR signal of the aromatic aldehyde near 9.85 appeared after acid

260

hydrolysis, strongly indicating the rupture of imine bonds. After the recycled dialdehyde structure was confirmed,

261

it was reused in the polyschiff vitrimer preparation with amine monomers, as shown in Figure 7a. As expected, a

262

flawless film was also obtained, showing the chemical recyclability of DAV monomer in polyschiff vitrimers.

1

1

1

263 264

Figure 7. (a) Chemical recycling of cross-linked polyimine films (conditions: 50 °C, pH =2, 24 h); (b) weight percentages of residuals after acid hydrolysis at

265

50 °C for 24 h

266

267 268

Figure 8. (a) FTIR and (b) 1H-NMR spectra of recycled DAV from acid hydrolysis (conditions: pH = 2, 50 °C, 24 h)

14

ACS Paragon Plus Environment

Page 14 of 21

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

ACS Sustainable Chemistry & Engineering

269

Conclusions

270

Bio-based polyschiff vitrimers were synthesized from vanillin and a commercial amine cross-linker. With the

271

introduction of dynamic imine covalent bonds into these crosslinked networks, the obtained bio-based vitrimers

272

show self-healing ability, hot-reprocessing properties, and chemical recyclability under acid hydrolysis. Even after

273

three hot-processing cycles, the tensile strength and elongation at break remained at the same level as those of

274

the original, sometimes even slightly higher. Importantly, dialdehyde monomers could be recycled from polyschiff

275

vitrimers under acid solution at 50 °C, for which structure is the same as that of the synthesized dialdehyde

276

monomer and could be used to regenerate polyschiff vitrimers. This work offers a significant opportunity for

277

designing self-healing and chemically recyclable polymer materials from lignin-based monomers.

278

Supporting Information

279

Figure S1. FTIR spectra of vanillin and DAV; Figure S2. NMR spectra of DAV monomer; Figure S3. FTIR spectra

280

of pristine polyschiff vitrimers; Figure S4. DSC results of pristine polyschiff vitrimers; Figure S5. TG results of

281

pristine polyschiff vitrimers; Figure S6. Strain-stress plot of pristine polyschiff vitrimers; Figure S7. FTIR spectra of

282

fractured sample 1 before and after 200 g weight hot-pressing; Table S1. Mechanical properties of pristine and

283

recycled polyschiff vitrimers using hot-processing.

284

Conflicts of interest

285

286

There are no conflicts to declare.

Acknowledgements

15

ACS Paragon Plus Environment

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

287

This work was supported by the foundation of Science Research Program from the Hubei Provincial

288

Department of Education (Grant No.Q20171602) and the Opening Project of State Key Laboratory of Polymer

289

Materials Engineering (Sichuan University) (Grant No. sklpme2017-4-14).

290

References

291

(1)

292

Langan, P.; Naskar, A. K.; Saddler, J. N.; Tschaplinski, T. J.; Tuskan, G. A.; Wyman, C. E. Lignin Valorization: Improving Lignin Processing in

293

the Biorefinery. Science. 2014, 344(6185), 1246843, DOI 10.1126/science.1246843.

294

(2)

295

Lignins. J. Am. Chem. Soc. 2018, 140(11), 4054-4061, DOI 10.1021/jacs.7b13620.

296

(3)

297

lignin-based aldehyde precursors. Eur. Polym. J. 2016, 74, 296-309, DOI org/10.1016/j.eurpolymj.2015.11.036.

298

(4)

299

DOI org/10.1016/j.progpolymsci.2013.11.004.

300

(5)

301

monomer synthesis. Green Chem. 2014, 16(4), 1987-1998, DOI 10.1039/C3GC42613K.

302

(6)

303

Eng. 2016, 4(1), 35-46, DOI 10.1021/acssuschemeng.5b01344.

304

(7)

305

Chem Rev. 2018, 118(2), 614-678, DOI 10.1021/acs.chemrev.7b00588.

306

(8)

307

Sustainable Chem. Eng. 2016, 4(11), 6082-6089, DOI 10.1021/acssuschemeng.6b01446.

Ragauskas, A. J.; Beckham, G. T.; Biddy, M. J.; Chandra, R.; Chen, F.; Davis, M. F.; Davison, B. H.; Dixon, R. A.; Gilna, P.; Keller, M.;

Gioia, C.; Lo Re, G.; Lawoko, M.; Berglund, L. Tunable Thermosetting Epoxies Based on Fractionated and Well-Characterized

Foyer, G.; Chanfi, B. H.; Boutevin, B.; Caillol, S.; David, G. New method for the synthesis of formaldehyde-free phenolic resins from

Laurichesse, S.; Avérous, L. Chemical modification of lignins: Towards biobased polymers. Prog. Polym. Sci. 2014, 39(7), 1266-1290,

Fache, M.; Darroman, E.; Besse, V.; Auvergne, R.; Caillol, S.; Boutevin, B. Vanillin, a promising biobased building-block for

Fache, M.; Boutevin, B.; Caillol, S. Vanillin Production from Lignin and Its Use as a Renewable Chemical. ACS Sustainable Chem.

Sun, Z.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K. Bright Side of Lignin Depolymerization: Toward New Platform Chemicals.

Zhao, S.; Abu-Omar, M. M. Renewable Epoxy Networks Derived from Lignin-Based Monomers: Effect of Cross-Linking Density. ACS

16

ACS Paragon Plus Environment

Page 16 of 21

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

ACS Sustainable Chemistry & Engineering

308

(9)

309

Lignin-Derived Phenols. ACS Sustainable Chem. Eng. 2017, 5(6), 5059-5066, DOI 10.1021/acssuschemeng.7b00440.

310

(10)

311

Resins Derived from Vanillyl Alcohol. ACS Sustainable Chem. Eng. 2016, 4(8), 4328-4339, DOI 10.1021/acssuschemeng.6b00835.

312

(11)

313

Epoxy Thermosets. ACS Sustainable Chem. Eng. 2018, 6(8), 11008-11017, DOI 10.1021/acssuschemeng.8b02419.

314

(12)

315

Preparation of UV curable Resins Together with Soybean Oil Derivatives. RSC Adv. 2016, 6(22), 17857-17866, DOI 10.1039/C6RA01420H.

316

(13)

317

J. 2015, 68, 526-535, DOI org/10.1016/j.eurpolymj.2015.03.048.

318

(14)

319

Lomonaco, D. Microwave-assisted solvent-free synthesis of novel benzoxazines: A faster and environmentally friendly route to the

320

development of bio-based thermosetting resins. J. Polym. Sci., Part A: Polym. Chem. 2017, 55(21), 3534-3544, DOI 10.1002/pola.28755.

321

(15)

322

DOI 10.1039/C3CS60044K.

323

(16)

324

2011, 334(6058), 965-968, DOI 10.1126/science.1212648.

325

(17)

326

3218(5), 1-9, DOI 10.1038/ncomms4218.

327

(18)

328

25(35), 4912-4917, DOI 10.1002/adma.201301513.

329

(19)

Zhao, S.; Abu-Omar, M. M. Synthesis of Renewable Thermoset Polymers through Successive Lignin Modification Using

Hernandez, E. D.; Bassett, A. W.; Sadler, J. M.; La Scala, J. J.; Stanzione, J. F. Synthesis and Characterization of Bio-based Epoxy

Savonnet, E.; Grau, E.; Grelier, S.; Defoort, B.; Cramail, H. Divanillin-Based Epoxy Precursors as DGEBA Substitutes for Biobased

Dai, J.; Jiang, Y.; Liu, X.; Wang, J.; Zhu, J. Synthesis of Eugenol-based Multifunctional Monomers via Thiol–ene Reaction and

Fache, M.; Viola, A.; Auvergne, R.; Boutevin, B.; Caillol, S. Biobased epoxy thermosets from vanillin-derived oligomers. Eur. Polym.

Oliveira Jéssica, R.; Kotzebue Lloyd, R. V.; Ribeiro Francisco, W. M.; Mota Beatriz, C.; Zampieri, D.; Mazzetto Selma, E.; Ishida, H.;

Jin, Y.; Yu, C.; Denman, R. J.; Zhang, W. Recent advances in dynamic covalent chemistry. Chem. Soc. Rev. 2013, 42(16), 6634-6654,

Montarnal, D.; Capelot, M.; Tournilhac, F.; Leibler, L. Silica-Like Malleable Materials from Permanent Organic Networks. Science.

Ying, H.; Zhang, Y.; Cheng, J. Dynamic urea bond for the design of reversible and self-healing polymers. Nat. Commun. 2014,

Yu, S.; Zhang, R.; Wu, Q.; Chen, T.; Sun, P. Bio-Inspired High-Performance and Recyclable Cross-Linked Polymers. Adv Mater. 2013,

Rekondo, A.; Martin, R.; Ruiz de Luzuriaga, A.; Cabanero, G.; Grande, H. J.; Odriozola, I. Catalyst-free room-temperature

17

ACS Paragon Plus Environment

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

330

self-healing elastomers based on aromatic disulfide metathesis. Mater. Horiz. 2014, 1(2), 237-240, DOI 10.1039/C3MH00061C.

331

(20)

332

Ester Bonds. J. Am. Chem. Soc. 2015, 137(20), 6492-6495, DOI 10.1021/jacs.5b03551.

333

(21)

334

Chem. Soc. 2017, 139(25), 8678-8684, DOI 10.1021/jacs.7b03967.

335

(22)

336

Transcarbamoylation. Angew. Chem., Int. Ed. 2016, 55(38), 11421-11425, DOI 10.1002/anie.201602847.

337

(23)

338

Recyclability. Macromolecules. 2017, 50(21), 8588-8597, DOI 10.1021/acs.macromol.7b01889.

339

(24)

340

for recoverable adhesives. Green Chem. 2018, 20(13), 2995-3000, DOI 10.1039/C8GC01299G.

341

(25)

Belowich, M. E.; Stoddart, J. F. Dynamic imine chemistry. Chem. Soc. Rev. 2012, 41(6), 2003-2024, DOI 10.1039/C2CS15305J.

342

(26)

Chao, A.; Negulescu, I.; Zhang, D. Dynamic Covalent Polymer Networks Based on Degenerative Imine Bond Exchange: Tuning the

343

Malleability and Self-Healing Properties by Solvent. Macromolecules. 2016, 49(17), 6277-6284, DOI 10.1021/acs.macromol.6b01443.

344

(27)

345

Acid (ELP–HA) Hybrid Hydrogels with Secondary Thermoresponsive Crosslinking for Injectable Stem Cell Delivery. Adv. Funct. Mater.

346

2017, 27(28), 1605609, DOI 10.1002/adfm.201605609.

347

(28)

348

Behavior. Adv. Funct. Mater. 2018, 28(14), 1706918, DOI 10.1002/adfm.201706918.

349

(29)

350

Thermoresponsive Triblock Copolymers as Potential Nanocarriers for “pH-Triggered” Drug Release. Macromolecules. 2011, 44(6),

351

1327-1334, DOI 10.1021/ma102804h.

Cromwell, O. R.; Chung, J.; Guan, Z. Malleable and Self-Healing Covalent Polymer Networks through Tunable Dynamic Boronic

Liu, W. X.; Chi, Z.; Zhang, H.; Ning, Z.; Yu, Z. X.; Jian, X. Oxime-Based and Catalyst-Free Dynamic Covalent Polyurethanes. J. Am.

Zheng, N.; Fang, Z.; Zou, W.; Zhao, Q.; Xie, T. Thermoset Shape-Memory Polyurethane with Intrinsic Plasticity Enabled by

Liu, T.; Hao, C.; Wang, L.; Li, Y.; Liu, W.; Xin, J.; Zhang, J. Eugenol-Derived Biobased Epoxy: Shape Memory, Repairing, and

Zhang, S.; Liu, T.; Hao, C.; Wang, L.; Han, J.; Liu, H.; Zhang, J. Preparation of a lignin-based vitrimer material and its potential use

Wang, H.; Zhu, D.; Paul, A.; Cai, L.; Enejder, A.; Yang, F.; Heilshorn Sarah, C. Covalently Adaptable Elastin-Like Protein–Hyaluronic

Ming, Z.; Fan, J.; Bao, C.; Xue, Y.; Lin, Q.; Zhu, L. Photogenerated Aldehydes for Protein Patterns on Hydrogels and Guidance of Cell

Xu, X.; Flores, J. D.; McCormick, C. L. Reversible Imine Shell Cross-Linked Micelles from Aqueous RAFT-Synthesized

18

ACS Paragon Plus Environment

Page 18 of 21

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

ACS Sustainable Chemistry & Engineering

352

(30)

353

pseudo target release behavior. Polym. Chem. 2018, 9(7), 878-884, DOI 10.1039/C7PY02108A.

354

(31)

355

Recyclability Enabled by Malleable Polyimine Networks. Adv Mater. 2016, 28(15), 2904-2909, DOI org/10.1002/adma.201505245 .

356

(32)

357

network polymer. Adv Mater. 2014, 26(23), 3938–3942, DOI org/10.1002/adma.201400317 .

358

(33)

359

Poly(dimethyl siloxane) Elastomer. Macromol. Rapid Commun. 2017, 38(15), 1700110, DOI 10.1002/marc.201700110.

360

(34)

361

imine bond. J. Polym. Sci., Part A: Polym. Chem. 2017, 55(12), 2011-2018, DOI 10.1002/pola.28577.

362

(35)

363

with Metal Ions. ISRN Polym. Sci. 2012, 2012, 1-5, DOI 10.5402/2012/532171.

364

(36)

365

coated magnetite nanoparticles for efficient removal of Pb(II) from water. J. Solid State Chem. 2016, 238, 311-320, DOI

366

org/10.1016/j.jssc.2016.04.003.

367

(37)

368

Polym. Sci., Part A: Polym. Chem. 2017, 55(10), 1790-1799, DOI 10.1002/pola.28544.

369

(38)

370

26(9), 1477-1485, DOI org/10.1002/adfm.201505368 .

371

(39)

372

Room Temperature Engineered by Transparent Elastomers. Adv Mater. 2018, 30(1), 1705145-1705141/1705148, DOI

373

10.1002/adma.201705145.

Tao, Y.; Liu, S.; Zhang, Y.; Chi, Z.; Xu, J. A pH-responsive polymer based on dynamic imine bonds as a drug delivery material with

Taynton, P.; Ni, H.; Zhu, C.; Yu, K.; Loob, S.; Jin, Y.; Qi, H. J.; Zhang, W. Repairable Woven Carbon Fiber Composites with Full

Taynton, P.; Yu, K.; Shoemaker, R. K.; Jin, Y.; Qi, H. J.; Zhang, W. Heat- or water-driven malleability in a highly recyclable covalent

Zhang, B.; Zhang, P.; Zhang, H.; Yan, C.; Zheng, Z.; Wu, B.; Yu, Y. A Transparent, Highly Stretchable, Autonomous Self-Healing

Zhang, H.; Wang, D.; Liu, W.; Li, P.; Liu, J.; Liu, C.; Zhang, J.; Zhao, N.; Xu, J. Recyclable polybutadiene elastomer based on dynamic

Amarasekara, A. S.; Razzaq, A. Vanillin-Based Polymers-part II: Synthesis of Schiff Base Polymers of Divanillin and Their Chelation

Culita, D. C.; Simonescu, C. M.; Patescu, R.-E.; Dragne, M.; Stanica, N.; Oprea, O. o-Vanillin functionalized mesoporous silica –

Ma, Z.; Wang, Y.; Zhu, J.; Yu, J.; Hu, Z. Bio-based epoxy vitrimers: Reprocessibility, controllable shape memory, and degradability. J.

Lyon, G. B.; Baranek, A.; Bowman, C. N. Scaffolded Thermally Remendable Hybrid Polymer Networks. Adv. Funct. Mater. 2016,

Kim, S.-M.; Jeon, H.; Shin, S.-H.; Park, S.-A.; Jegal, J.; Hwang, S. Y.; Oh, D. X.; Park, J. Superior Toughness and Fast Self-Healing at

19

ACS Paragon Plus Environment

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

374

(40)

375

2,2,6,6-Tetramethylpiperidine-1-sulfanyl (TEMPS) Radicals. Angew. Chem., Int. Ed. 2017, 56(8), 2016-2021, DOI 10.1002/anie.201611049.

376

(41)

377

thermoplastics through dioxaborolane metathesis. Science. 2017, 356(6333), 62-65, DOI 10.1126/science.aah5281.

378

(42)

379

30-38, DOI 10.1039/C5SC02223A.

380

(43)

381

Polym. Sci., Part A: Polym. Chem. 2017, 55(23), 3826-3831, DOI 10.1002/pola.28856.

Takahashi, A.; Goseki, R.; Otsuka, H. Thermally Adjustable Dynamic Disulfide Linkages Mediated by Highly Air-Stable

Röttger, M.; Domenech, T.; Van, d. W. R.; Breuillac, A.; Nicolaÿ, R.; Leibler, L. High-performance vitrimers from commodity

Denissen, W.; Winne, J. M.; Du Prez, F. E. Vitrimers: permanent organic networks with glass-like fluidity. Chem. Sci. 2016, 7(1),

Collins, J.; Xiao, Z.; Connal Luke, A. Tunable degradation of polyethylene glycol-like polymers based on imine and oxime bonds. J.

382 383

20

ACS Paragon Plus Environment

Page 20 of 21

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

ACS Sustainable Chemistry & Engineering

384

TOC:

385 386

Thermally reprocessable and acid-degradable vanillin-derived vitrimers were prepared, showing a total chemical

387

recyclability of bio-resource based materials.

388

21

ACS Paragon Plus Environment