Potential of Organosolv Lignin Based Materials in Pressure Sensitive

Jul 11, 2019 - In this study, a commercial superplasticizer, a polycarboxylate polyether (PCE) is blended with a beech-based organosolv lignin, and th...
0 downloads 0 Views 818KB Size
Subscriber access provided by Nottingham Trent University

Article

Potential of Organosolv Lignin Based Materials in Pressure Sensitive Adhesive Applications Gopakumar Sivasankarapillai, Elahe Eslami, and Marie-Pierre G. Laborie ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/acssuschemeng.9b01670 • Publication Date (Web): 11 Jul 2019 Downloaded from pubs.acs.org on July 18, 2019

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 27 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

Potential of Organosolv Lignin Based Materials in Pressure Sensitive Adhesive

2

Applications

3

Gopakumar Sivasankarapillai1,2, Elahe Eslami1, Marie-Pierre Laborie1,2

4

1

Freiburg Material Research Center, Stefan-Meier-Strasse 21, D-79104, Albert-Ludwig- University

5

of Freiburg, Freiburg, Germany

6

2

7

6, D-79085, Albert-Ludwig- University of Freiburg, Freiburg, Germany

8

E-mail: [email protected]

9

ABSTRACT

Chair of Forest Biomaterials, Faculty of Environment and Natural Resources, Werthmannstrasse

10

Blending lignin with other polymers has long been utilized as a potent approach to explore novel

11

applications for lignin. Yet, combinations of organosolv lignin (OSL) with superplasticizer have been

12

hardly explored as a potential to design novel material properties. In this study, a commercial

13

superplasticizer, a polycarboxylate polyether (PCE) is blended with a beech-based organosolv lignin

14

and their miscibility and adhesive properties are studied over the entire compositional range. All

15

blend compositions exhibit a single glass transition temperature (Tg) that follows the Kwei model of

16

miscible polymer blends, while a thermal stabilization effect is evidenced in blends containing up to

17

50% lignin. Fourier transform infrared spectroscopy (FTIR) investigations reveal specific

18

interactions between functional groups in PCE and lignin OH groups as molecular basis for

19

miscibility. The rheological properties, tackiness and adhesive peel strength suggest that these

20

blends have potential as base formulation for pressure sensitive adhesive applications, albeit

21

needing some fine-tuning.

22

KEYWORDS: Organosolv Lignin, Polycarboxylate Polyether, Pressure Sensitive Adhesive, Blend

23

Miscibility, Thermal & Rheological Properties

24

INTRODUCTION 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

Page 2 of 27

25

As a by-product of the pulp and paper industry and the second most abundant biopolymer on

26

earth, lignin has long been considered a valuable biomacromolecule for material development. In

27

particular, lignin-based polymer blends have long garnered attention from scientists as testified by the

28

number and the timeframe of reviews on this topic. 1–7

29

Recognizing lignin potential in the late 80’s, Glasser and his group pioneered blending approaches

30

of lignin and its derivatives with a wide range of polymers including biodegradable polymers such

31

as cellulose acetate butyrate (CAB), hydroxypropyl cellulose (HPC), Polyhydroxybutyrate (PHB),

32

Polycaprolactone (PCL), lignin esters, starch, and petroleum-based polymers such as PVC. 8 With

33

the application as carbon fibers in view, Kadla and coworkers later studied the miscibility between

34

hardwood Kraft lignin and synthetic polymers 9-12 particularly investigating blend properties and

35

intermolecular interactions in hardwood kraft lignin and synthetic polymers such poly(ethylene

36

oxide) (PEO), poly(vinyl alcohol) (PVA), poly(ethylene terephthalate) (PET) and poly(propylene)

37

(PP). Also prevalent in the literature since the late 80s is the utilization of lignin in adhesive

38

applications13 and particularly in structural adhesives. Here lignin can be used as phenol substitute

39

in wood adhesives while lowering their carbon footprint and cost.14

40

More recently biomass-derived polymers including industrial lignin have also been considered for

41

pressure sensitive adhesive (PSA) applications.15-16 Pressure sensitive adhesive (PSAs) are soft

42

viscoelastic solids that adhere onto the substrate on application of a slight pressure in a very short

43

time18. PSA require three important properties: all PSAs require some degree of stickiness or tack;

44

removable PSA require a controlled peel force and adhesive failure for debonding while permanent

45

PSAs must exhibit minimal creep.17 To meet these requirements, PSA are formulated with multiple

46

components, with at least one component operating well above its Tg. For example, water-based

47

PSAs consist of dispersions of a high Tg component and a low Tg component, whose composition

48

determines the critical properties of the PSA i.e. the tack, peel strength and shear strength.

2 ACS Paragon Plus Environment

Page 3 of 27 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

49

In this work we hypothesize that blends of high Tg lignin with a very low Tg polymer might well

50

perform as a PSA, opening new applications for lignin. We choose to investigate the suitability of

51

aqueous blends of organosolv lignin, a high Tg macromonomer, with a superplastizer as the low

52

Tg component, for PSA application. We select organosolv lignin because in contrast to industrial

53

lignin, organosolv lignin exhibits higher purity (no sulfur) and homogeneity, lower molecular weight

54

and glass transition temperature.18, 19 We further select as the low Tg component a

55

polycarboxylate polyether (PCE), a commercial superplasticizer commonly used for admixtures in

56

concrete. Besides its very low Tg, its functional groups suggest that it is able to interact with lignin

57

(Scheme 1). In concrete, lignin has for example been repeatedly used for the modification of the

58

aliphatic superplasticizer, 20-22 Intermolecular interactions and miscibility among the components

59

are in fact critical to the PSA performance as it influences its viscoelastic properties. 23-25

60

In this paper we first report on the miscibility and intermolecular interactions in the organosolv

61

lignin/PCE pair. We then examine the rheological behavior, tack, shear and peel strength of blends

62

of varying composition, to assess their suitability for PSA applications.

63

EXPERIMENTAL SECTION

64

Raw Materials:

65

Polycarboxylate Polyether (PCE) in aqueous solution/sodium salt (ETHACRYL.HF) was received

66

from Arkema (France) and used without further modification or purification (Scheme 1).

67

Commercial crepe paper latex-impregnated tape, (Würth), was used as the substrate for testing

68

adhesive properties. Beech Organosolv lignin (OSL) was received from Fraunhofer CBP,

69

Germany.

70

Raw Materials Characterization:

71

The molecular weight of the PCE was analyzed on a GPC SECurity 1200 system (PSS-Polymer

72

Standards Service, USA) (Table 1). Polystyrene narrow standards were used for establishing the 3 ACS Paragon Plus Environment

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

Page 4 of 27

73

calibration curve. For lignin, the total hydroxyl group content was determined with 31P NMR

74

(Table1) following published procedure.26

75

76 77

Scheme 1: Hypothetical partial structure of beech wood organosolv lignin 27 (above) and commercial

78

PCE (bottom).

79

Table 1. Molecular weight, glass transition temperature and 31PNMR data of raw materials 31P

NMR (mmol/g OSL

Raw

Tg Mw(g/mol)

Polydispersity

Aliphatic

pH

Aromatic (oC)

material

OSL(Beech)

Solid

4300*

3.9*

46535

1.69

OH

OH

2.43

2.39

content

114

-

-60

3.7*

-

PCE(in _

_

40*

aq.solution) 80

*Data received from the manufacturer

81

In brief, an exact amount of 25–30 mg of the OSL samples was diluted in 400 µL CDCl3/pyridine

82

(1:1.6) ,150 µL of a solution of chromium(III) acetylacetonate (3.6 mg/mL) as relaxation agent and 4 ACS Paragon Plus Environment

Page 5 of 27 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

83

cyclohexanol (4.0 mg/mL) as an internal standard in CDCl3/pyridine (1:1.6) were added and the

84

solution was stirred for 5 min. 2-Chloro-4,4,5,5-tetramethyl-1,2,3-dioxaphospholane (TMDP, 70 µL)

85

was added and the solution was transferred into an NMR tube for analyzing in a Bruker 300 MHz

86

spectrometer with 128 scans and a delay time of 15 sec. The chemical shifts relative to the

87

reaction product of TMDP with water at 132.2 ppm are assigned to the functional groups at δ =

88

150.0–145.5 (aliphatic-OH), 145.5–144.7 (cyclohexanol), 144.7–136.6 (phenolic-OH), 136.6–133.6

89

(carboxylic acids) ppm.26 OSL and the PCE display very distinct Tg and the PCE appears as a

90

rather low molecular weight polymer.

91

Blends Preparation

92

OSL and the PCE aqueous solution were mixed thoroughly in a beaker at 100 0C for 15 min with

93

constant stirring on a hot plate in PCE:OSL w/w blend compositions of 100:0, 80:20, 70:30, 60:40,

94

50:50, 40:60, 20:80, 0:100 (5g solid content batch). Then all the samples were dried in a vacuum

95

oven for 1h at 1500C. The moisture content of the dried blends was assessed with

96

thermogravimetric analysis and did not exceed 3%. The dried blends were subjected to various

97

analyses to determine miscibility, nature of molecular interactions, rheological behavior and

98

adhesive performance of PSA.

99

Thermal and Spectroscopic Characterization of PCE:OSL Blends

100

Differential scanning calorimetry (DSC): about 8-10 mg of sample was placed in a pin-holed and

101

sealed 30 μl Aluminum pan (3 replicates) and analyzed on a DSC8500 Pyris 1 (Perkin Elmer,

102

USA) at 100C/min under nitrogen atmosphere. The PCE sample was heated from 30 to 100°C at

103

20°C/min in the first scan, then the sample was cooled to -90 at 20°C/min to minimize the enthalpy

104

relaxation in the second heating scan. After two minutes of holding time, the sample was reheated

105

to 150°C at 10°C /min in order to determine the glass transition temperature (Tg). The starting (-90

5 ACS Paragon Plus Environment

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

Page 6 of 27

106

to -30°C) and the end temperatures (100°C to 250°C) were changed depending on the blend

107

composition.

108

Thermogravimetric analysis (TGA): about 5 mg of samples placed in thermogravimetric analyzer,

109

Pyris 1 (Perkin Elmer, USA) and heated to 9000C at 10 0C /min under nitrogen stream. The TGA

110

was used to evaluate the water content and the thermal stability of the PCE:OSL blends by virtue

111

of the thermal decomposition temperature, and the temperatures at 5% and 10% mass loss,

112

respectively.

113

Fourier transform infrared spectroscopy (FTIR): spectra were collected on a FTIR spectrometer 65

114

(Perkin Elmer, USA) operating in the attenuated total reflection (ATR) mode (ZnSe crystal). Each

115

spectrum was taken as an average of 32 scans at a resolution of 4 cm-1. For the 30wt% OSL

116

content blend, the theoretical spectrum was mathematically constructed according to the rule of

117

mixture and compared with the experimental spectrum in order to delineate specific interactions in

118

the blends.

119

Rheological Properties of PCE:OSL Blends

120

Rheological measurements were carried out on a HAAKE Mars-II rheometer with Modular

121

Advanced Rheometer System. The oscillatory experiments of the samples with 20 mm diameter

122

and 1.0-0.5 mm thickness were performed on the parallel-plate fixture. The angular frequency

123

ranged from 0.01 to 100 Hz in the linear viscoelastic zone (5%) at Tg+700C temperatures.

124

Adhesive Performance of PCE:OSL Blends

125

The blends were evaluated for performance in relation to a possible use as a PSA. Namely, tack,

126

peel adhesion and shear strength were determined for 0-40wt% OSL compositions.

127

The tack values were determined with the rolling ball tack method as the distance traveled by the

128

ball on an adhesive layer before stopping in accordance with ASTM D 3121 standard.28 In 6 ACS Paragon Plus Environment

Page 7 of 27 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

129

general, rolling ball travel distances are divided to three zones: i) 0 to 100 mm is good tack zone; ii)

130

100 to 200 mm is medium tack zone and iii) 200 to 300 mm is low tack zone.29 The 180o peel test

131

was performed in accordance with the Pressure Sensitive Tape Council standard for 180o peel test

132

(PSTC 101, test A).30 The universal tensile tester, Inspect mini Hegewald & Perschke was used for

133

running 180o peel test. A 24 ± 0.5mm wide and 300mm long strip of the PCE:OSL coated film was

134

prepared. Tests were performed on 120 micrometer (μm) thick films with a 20 μm coating of

135

adhesive. The thickness of adhesive was measured by “Mitutoyo” Thickness Gages (Accuracy: ± 2

136

μm). After preparation of the films, they were adhered on the test panel. The test panel is a 50 by

137

125mm and 1.1mm thick stainless-steel panel in accordance with specification of ASTM A 666.31

138

The stainless-steel panel was clamped into the bottom jaw of the adhesion tester, the tape doubled

139

back at an angle of 180o and clamped into the upper jaw. Tensile testing machine was operated at

140

5.0±0.2 mm/s, at room temperature (23±-2OC and 50±5% relative humidity). The motion of

141

movable jaw causes the mechanically peeling of the film. The average force per unit width

142

obtained during peeling is considered as the adhesion value. Five different specimens were tested

143

for each composition.

144

The lap joint shear strength test was carried using tensile testing machine (Inspeck mini Hegwald

145

& Perschke) to determine the shear strength of PCE:OSL blends coated on glass strips according

146

to ASTM D1002. Specimen size for lap shear specimen was 25.4 mm wide, 76 mm long with an

147

overlap of 12.7 mm. The blends were coated in thin layers (25 g/ m2 and 20±5 μm) and

148

conditioned at around 23OC with 50± 5% relative humidity before the test. The testing distance was

149

50 mm. The loading applied was 80 to 100 kg/ cm2 of the shear area per min, and pulled at

150

1.3mm/min. until rupture occurs.

151

Finally, Shear resistance was investigated by examining the adhesive strength of the PCE:OSL

152

blends under a constant load according to PSTC-107 testing procedure.32 A 20mg of PCE:OSL

153

blend was coated on a PET silicon tape and the coated strip was placed on a horizontally mounted 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

Page 8 of 27

154

stainless steel plate with a contact area of 25mm x 25mm. A 100g weight was attached to the free

155

end of the PET strip with a metal clamp. The experiment was started by recording the time taken to

156

pull the adhered tape away from the stainless steel plate. The experiment was repeated with 0-40

157

wt% of OSL samples.

158

RESULTS AND DISCUSSION

159

Thermal Properties and Blend Miscibility

160

Miscible blends on a 10-20 nm ranges display one single glass transition temperature (Tg),

161

whereas immiscible blends exhibit the two glass transition temperatures of the parent polymers.33

162

Figure 1 shows the DSC curves from the second run of the PCE:OSL blends prepared by hot

163

mixing. An endothermic step is observed for all the samples and provides a signature of their Tg

164

(Figure 1). For vacuum-dried OSL and PCE, the Tg appears at 1140C and -610C, respectively.

165

The PCE:OSL blends also exhibit a single Tg, which varies between the values for the parent

166

polymers, depending on the blend composition, suggesting complete miscibility.34 Various models

167

have been proposed for predicting the composition dependence of Tg in miscible polymer blends,

168

shedding further light on the possible nature of the interactions. In particular, the Fox equation,

169

Gordon–Taylor (GT) and Kwei equations are commonly used for miscible polymer blends.35-37

170

(1)

Fox,

171

(2)

GT, (Tg) =

172

(3)

𝐾𝑤𝑒𝑖, ( 𝑇𝑔 ) =

173

Here, Tg is the glass transition temperature of the blend, Tg1, Tg2, w1, w2 are the glass transition

174

temperatures and weight fractions of the parent polymers, respectively. In the Gordon-Taylor and

1 w1 w2 = + Tg Tg1 Tg 2 w1 Tg1 + kw2 Tg2 w1 + kw2 𝑤1 𝑇𝑔1 + 𝑘𝑤2 𝑇𝑔2 + 𝑞𝑤1 𝑤2 𝑤1 + 𝑘𝑤2

8 ACS Paragon Plus Environment

Page 9 of 27

175

Kwei model, k and q are adjustable parameters reflecting intermolecular interactions. Normally the

176

q parameter corresponds to the strength of hydrogen bonding.38, 39

120

b) 80

Tg (0C)

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

Expt Fox GT Kwei

40 0 -40 -80 0

177

20

40

60

80

100

OSL wt%

178

Figure 1: DSC curves (left), and glass transition temperatures of PCE:OSL blends Vs OSL wt%

179

with experimental data points and theoretical predictions from the Fox, Gordon-Taylor and Kwei

180

models (right).

181

In Figure 1b, all three models globally describe the general trend in Tg with composition with the

182

Fox model resulting in the lowest goodness of fit (R = 0.96). The Kwei model provides the best fit

183

(R=0.99) to the experimental Tgs of blends with k =1.52±0.5 and q = -0.01, followed by the GT

184

model with kGT value of 2.35 (R= 0.98). The fact that the GT and Kwei model provides the best fit

185

to the data compared to the Fox equation9,37 suggests specific interactions between the parent

186

polymers. The value of k in the Kwei model indicate that OSL contribution to the final TgKwei is

187

proportionally smaller than its mass fraction.37 As a result, the curve obtained by the TgKwei shows a

188

negative shift (U-shaped) from the simple weight-averaged curve. The parameter q portrays the

189

balance between the breaking of self-associations and the forming of inter-associations between

190

the two parent polymers.39, 40 The q value close to zero implies that there is a good balance

191

between the breaking of intra-molecular geometry of OSL and the formation of intermolecular

192

interactions between PCE and OSL. Kadla et al investigated the supramacromolecular lignin 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 27

193

complexes break up through specific intermolecular interactions between polyethylene oxide

194

(PEO) and individual Kraft lignin components and showed that increasing PEO incorporation

195

disrupts the supramacromolecular structure of lignin.12 Likewise, the value of q in the present

196

PCE:OSL blend suggests that the dissociation of lignin supramolecular structure nullified the effect

197

on Tg due to the formation of supramacromolecular structure of PCE:OSL blends through H-

198

bonds.

199

Nature of intermolecular Interactions based on Fourier transform infrared (FTIR)

200

spectroscopy.

201

FTIR analysis might help shed light on the existence and nature of interactions in miscible polymer

202

blends.12,39,40 FTIR spectra of the PCE:OSL blends show significant and systematic changes upon

203

addition of OSL to PCE, in particular in the OH, C=O stretching region and in the C-O-C region

204

(Figure 2). With increasing OSL content, the hydroxyl band of OSL and PCE merge towards one

205

another, as might be expected from a simple mixture effect (Figure 2b). In addition, a shoulder

206

band appears around 3290 cm-1 in the experimental spectrum. This new band is not present in the

207

theoretical spectrum obtained by mathematical superposition of the parent spectra in the 30wt%

208

OSL blend (Figure S3). It thus suggests the appearance of additional hydrogen bonds in the

209

blends. The same trend is observed in the C-O-C region of PCE at 1096 cm-1 shifts to the lower

210

region, reaching for example 1089 cm-1 in blends with 30wt% OSL (Figure 2a), even farther away

211

from the lignin C-O-C vibration at ca. 1118 cm-1. Again, this differs from the theoretical spectrum

212

(Figure S3) where the band center at 1094 cm-1. Finally, the carbonyl vibrations of OSL (1732 cm-

213

1

214

simply obeying the law of mixture. The appearance of new OH stretching vibrations and the shifts

215

in C-O-C band center which do not simply obey a rule of mixture reveal the presence of specific

216

intermolecular interactions. Namely, the polyether graft chains of the PCE could be involved in H-

217

bonds with the lignin OH groups. The systematic trend in band center shifts for C-O-C vibrations

) and PCE (1725 cm-1) also shift in the blends, albeit in proportion to the component ratios, thus

10 ACS Paragon Plus Environment

Page 11 of 27 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

218

(Figure 3) upon OSL addition confirms that specific interactions involving these functional groups

219

are taking place. All the samples show changes in ether and in carbonyl regions with broadened

220

bands at lower wavenumber (1729 cm-1, Figure S4) probably due to the presence of different

221

carbonyl groups in different environments from both blend components.

222

To assess whether esterification did take place in the blends, a set of blends were prepared at

223

various temperatures and further characterized by FTIR (Figure S5). Blending at higher

224

temperatures, viz. at 170 OC and 190 OC revealed a new shoulder around 1759 cm-1, which is

225

characteristic of ester carbonyls.41 In the normal preparation conditions of the blends, this shoulder is not

226

detected, suggesting that esterification does not take place . Instead, the observed band changes

227

confirm the hypothesis that H-bonding takes place between the hydroxyl groups of OSL and the

228

carboxylic and ether functionalities of the PCE.

229 230

Figure 2: FTIR spectra of PCE:OSL blends and its components: PCE:OSL with 30wt%OSL),

231

70wt% OSL, PCE and OSL, FTIR spectra of (a) C-O-C region (b), OH stretching region.

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

Page 12 of 27

232 233

Figure 3: The relationship of absorbance of OSL and PCE based ether groups (band center

234

position) against wt% of OSL (left) and hypothetical architecture of supramolecular aggregates of

235

PCE:OSL blend (right)

236

Altogether the thermal analysis and vibrational spectroscopy provide strong evidence for miscibility

237

across the entire compositional range of PCE and OSL arising from specific H-bonding between

238

the lignin OH groups and the PCE component. As lignin /PEO miscibility is clearly established it is

239

no surprise that PCE, which comprises PEO branches exhibit also strong miscibility with lignin.12

240

Rheological Properties

241

Rotational rheological properties of the PCE:OSL were investigated using a parallel plate set up for

242

PCE:OSL blends with up to 50wt% OSL content. For all tested PCE:OSL blends, the storage

243

modulus (G’) is lower than the loss modulus (G”) in the entire range of frequencies (ω) tested

244

(Figure 4), which is indicative of a viscous behavior.42 This reveals that there is no chemical

245

network developing in the blends. Likewise, no crossover of G″ and G′ can be observed at room

246

temperature over the entire frequency range (1-100 rad/ ω) also supporting that gel formation does

247

not occur in all blend compositions at room temperature. Finally, G’ and G” are found to increase 12 ACS Paragon Plus Environment

Page 13 of 27 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

248

with increasing OSL content. Yang and Chang suggested that the viscoelastic properties of a

249

polymeric system as obtained by rheology provide insight as to its potential performance as a

250

pressure sensitive adhesive.43 At a frequency of 1 Hz, viz., in the bonding region the polymeric

251

systems should satisfy the Dahlquist criterion, which states that G’ had to be less than about 0.1

252

MPa before any adhesive tack was observed. 44. This criterion predetermines whether the material

253

has the required tackiness in the bonding region. This is clearly the case for all PCE:OSL

254

formulations tested including the neat PCE. In turn, the 180 O peel strength of a PSA can be

255

deduced by observing viscoelastic properties at a frequency of circa 435 rad/s. To accede to this

256

region, master curves were created for the blend systems by virtue of Time Temperature

257

Superposition (TTS) using Tref =30 OC and isothermal data acquired at 10 OC intervals between -20

258

O

259

the entire frequency and temperature ranges used and yielded smooth master curves and shift

260

factor curves (Figure 4, right). However, when examining the master curve of a single blend, for

261

example the 40wt% OSL compositions over a broad range of temperature, a crossover frequency

262

for G’ and G’’ clearly reveals the existence of a gel point close to the debonding frequency. In other

263

words, with increasing temperature the blends are able to form a physically crosslinked network

264

adopting a solid-like behavior.

C to 70 OC. The TTS and the William Landel Ferry equation appeared valid for all samples over

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

Page 14 of 27

265

Figure 4. Dynamic modulus (G’ &G’’) against the angular frequency of PCE:OSL blends with

266

different OSLwt% at Tg+700C (left), and master curve(right) for the 40wt% OSL blend showing a

267

gel point at ω = 103 rad/s. (Dahlquist line G’< 3. 105 Pa)

268

The ratio of the loss tangent to the storage modulus (tan δ /G’) provides information on the ability

269

of the adhesive to release energy at the debonding frequency.45 For example a value of 5 MPa-1

270

and 10 MPa-1 is recommended for stainless steel and polyethylene, respectively. For the neat

271

PCE, the tan δ/G' value is ca. 2.7 MPa-1 [Figure 5]. With initial addition of OSL this value increases

272

to ca. 4 MPa-1 and 3 MPa-1 for 10 wt% and 20wt% OSL respectively, and then decreases again

273

upon further addition. In the case of these two compositions, blending with lignin increases energy

274

dissipation mechanisms, possibly as a result of the supramolecular blend assembly by multiple

275

hydrogen bonds.46,47 While small amounts of lignin and thus the occurrence of a PCE:OSL

276

supramolecular assembly thus contribute to the energy release ability of the system, it remains

277

insufficient to provide adequate dissipation of energy at the debonding frequency. In other words,

278

the blends maintain a strong elastic character such that debonding on steel surfaces may occur by

279

interfacial cracks. Further tuning of the system is therefore required to improve energy dissipation

280

capacity.

281 14 ACS Paragon Plus Environment

Page 15 of 27 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

282 283

Figure 5: The ratio tan δ /G’ as a function of frequency and the tan δ /G' for PCE:OSL blends with different amounts of OSL measured @ 23±2 OC and 1 Hz (inset figure).

284

The viscoelastic data can be further utilized to determine the usable window of a PSA.18,48 In this

285

analysis materials falling into quadrant 1 are typically not suited for PSA applications. Quadrant 2

286

relates to PSA which necessitate high shear strength. Quadrant 3 applies to PSA with low peel

287

strengths and therefore that are easily removable. Quadrant 4 is the region of low G‘ and G ‘‘ and

288

stands for quick-stick PSA. In Figure 6, it is apparent that the neat PCE does not fit in the general

289

PSA window. In contrast, adding OSL to PCE shifts the viscoelastic window closer to the general

290

PSA. In particular the 20 and 30 wt% OSL content blends center on Quadrant 3, revealing possible

291

application as removable PSA. The 40 wt% OSL blend centers around Quadrant 2, suggesting

292

potential application as high cohesive strength PSAs. Regardless of the special attributes of the

293

lignin blend as PSA, it is clear that blend composition is a useful tool to tune the specialty of the

294

PSA from a removable PSA to a high shear PSA.

295 296

Figure 6. Viscoelastic window of PCE:OSL blends with different OSL wt% @ 23±2OC.

297

Adhesive Properties of PCE:OSL Blends

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

Page 16 of 27

298

The rheological characterizations on the blends have suggested that the PCE:OSL blends might

299

well function as PSA. Creton defines 3 properties that PSAs require.18 “First, they require some

300

tackiness in order to form good bonds, second they need a controlled peel force as a function of

301

peel velocity and precise control of the residue left, finally they must exhibit minimum creep”. To

302

further assess this possibility, adhesive tackiness, 180O peel strength and shear strength were

303

tested for 0-40wt% OSL blend compositions. Turning first to the tackiness, it is apparent that

304

adding OSL up to ca. 20wt% improves the material tackiness, at which point further addition

305

slightly lowers tackiness (Figure 7). A tackiness level similar to that of a commercial PSA (coating

306

thickness 17± 2 μm, similar to that of the PCE:OSL blends at 20± 2μm) could be obtained by the

307

addition of up to 20wt% OSL. The 30wt% OSL content blend maintains acceptable tackiness in

308

contrast to the 40wt% OSL content blend, which completely loses its tackiness. Tackiness is

309

further evidenced by the ability of the material to form fibrils upon extension and detachment in the

310

rheometer (Figure 8, right).

311 312

Figure 7:Tack, peel and shear strength measured @ 23±2OC for the PCE:OSL blends

16 ACS Paragon Plus Environment

Page 17 of 27 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

313 314

Figure 8: Shear strength of various composition of PCE:OSL blends @ 23±2OC(left) and a

315

PCE:OSL blend in rheometer set up (right)

316

The 180O peel test reveals that the neat PCE has a low peel strength (0.05 N/cm). According to

317

Lee and Shin 49, excellently removable PSA exhibit a peel strength below ca 0.5 N/cm. However,

318

as evidenced by the viscoelastic analysis, the neat PCE does not fall into the usable window of a

319

PSA. Increasing OSL loading up to 30wt% increases peel strength more than 10 fold reaching

320

around 7 N/cm for the 30% OSL content blend and 2.2 N/cm for the 20wt% OSL content blend

321

(Figure 7). This positions the 20wt% OSL content blend in the category of the removable and

322

repositionable or semi-removable PSA such as scotch tape, while the 30wt% OSL content blend

323

pertains to the permanent PSA such as duct tape, in respect to their peel load.49 Therefore, these

324

two blends have peel strengths well in line with the performance of commercial removable and/or

325

repositionable PSAs. Again, lignin content is clearly revealed as a tool to engineer the peel

326

strength of the resulting adhesive system.

327

The shear test was done to measure the shear resistance of the PCE:OSL blends. The neat PCE

328

showed no strength at the testing condition as in fact, the PCE coated strip fell from the steel plate

329

as soon as the weight was hanged. Figure 8 displays the stress-strain curve for the shear test of

330

the different systems and confirms a low shear strength for all PCE:OSL blends with less than 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

Page 18 of 27

331

30wt% OSL content. It also clearly reveals that increasing OSL content up to 40wt% delivers a

332

shear strength that is similar to that of a commercial PSA (Würth tape). With its glassy state and

333

its ability to form hydrogen-bonded networks with PCE, lignin contributes resistance to shear to the

334

system. The reason for the drastic increase in shear resistance at exactly 40% is however

335

unknown. Perhaps a change in the blend microstructure occurs, but this was not specifically

336

monitored. The shear resistance was also measured by examining the adhesive strength of the

337

PCE:OSL blends under a constant load according to PSTC-107 testing procedure.32 Adding OSL

338

to the PCE here again improved the shear resistance of the system, albeit it remained low.

339

PCE:OSL blends with 10, 20 and 30wt% of OSL could withstand the load for 1, 4 and 9 min

340

respectively. The test with 40wt% OSL content blend failed due the lack of tackiness. The obtained

341

time from 0-30wt% OSL blends remains low in comparison to other reports in the literature with

342

other biobased PSA systems and commercial PSAs but approaches the values for post-it

343

tapes.17,49,50.

344

Significant differences in tack, peel resistance and shear strength due to the presence of OSL

345

were detected. Interestingly, increasing the volume fraction of OSL to 30 wt% improves all the

346

critical PSA properties, except the tackiness which is only very moderately lowered. The

347

microstructural origin of the performance of the PCE:OSL is difficult to pinpoint with this study.

348

However, a few statements can be safely made based on prior literature.46,47,51-54 The

349

supramolecular assembly of lignin and PCE via hydrogen bonding likely provides some additional

350

energy dissipation mechanisms and tackiness as observed for the 10 and 20wt% OSL blends.

351

The ability to form physically crosslinked polymer networks with the glassy lignin moieties further

352

provides shear resistance to the assembly. The higher the physical crosslinking density, as is

353

expected with increasing lignin content, the higher the shear resistance. It is thus the

354

supramolecular assembly via hydrogen bonding itself that would be responsible for the

355

improvement in tackiness, peel strength and shear resistance as lignin is incorporated in moderate

356

amounts into the PCE. Above a critical lignin content of between 20 and 30 wt%, lignin plays 18 ACS Paragon Plus Environment

Page 19 of 27 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

357

mainly the role of a glassy filler which enhance shear resistance and peel strengths but the minor

358

supramolecular assembly of PCE:OSL systems no longer suffices to enhance energy dissipation

359

capability. This particularly clear for the 40 wt % OSL blend, which exhibits maximum shear and

360

peel strength and minimum tan/G’ and tackiness. In contrast, the PCE with its Tg below room

361

temperature functions as sticky branches.

362

Considering all the performances together, it stands out that PCE:OSL systems are possible base

363

formulations for PSA applications. Further fine tuning of the formulations is needed to completely

364

optimize the systems for PSA application, especially regarding the energy dissipation ability.

365

CONCLUSION

366

Organosolv lignin-based blends have been produced using a commercial PCE. A single Tg well

367

modeled by the Kwei analysis for the PCE:OSL blends alongside FTIR analyses evidenced

368

PCE:OSL miscibility and the existence of specific interactions between the OSL hydroxyl and the

369

proton accepting sites of PCE. Miscibility between the two parent polymers improved the thermal

370

stability and viscoelastic response of the blend in view of their possible utilization as pressure

371

sensitive adhesives. In particular, the blends up to 30wt% OSL content exhibited favorable tack,

372

peel and shear properties that could be tuned by the OSL content to fit specialty application from

373

removable/repositionable or semi removable PSA. The blends supramolecular assembly by

374

hydrogen bonding is proposed to be in part responsible for additional energy dissipation

375

mechanisms in the blends, while the glassy lignin and the rubbery PCE are mostly responsible for

376

shear strength and tackiness, respectively. Nonetheless additional fine tuning, possibly including

377

mild covalent crosslinking is needed to completely match the ideal performance of any category of

378

PSA.

379

ASSOCIATED CONTENT

380

Supporting Information 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

Page 20 of 27

381

Thermogravimetric analysis, Mathematically constructed and experimentally observed FTIR

382

spectra 30wt%OSL blend; FTIR spectra of carbonyl stretching region of blends, FTIR spectra of

383

selected PCE:OSL blends prepared at different temperatures.

384

AUTHOR INFORMATION

385

Corresponding author

386

* Phone: +49 (0)761 203 4759. Fax:

387

E-mail: [email protected]

388

Notes

389

Conflicts of Interest: The authors declare no conflict of interest

390

ACKNOWLEDGMENTS:

391

This work was partially funded by the European Union’s Seventh Framework Program, FP7-NMP,

392

and Project number 604215 (CARBOPREC). The authors wish to thank Arkema (France) and

393

Fraunhofer CBP, Germany for providing Polycarboxylate Polyether (ETHACRYL.HF) and

394

(Organosolv lignin) respectively. The authors would like to thank Elke Stibal for her technical

395

assistance.

+49 761 203 37 63

396 397

REFERENCES

398 399

1. Wang, J.; Manley, R.St.J.; D. Feldman, Synthetic polymer-lignin copolymers and blends,

400 401 402

Prog. Polym. Sci. 1992, 17, 611-646, DOI:10.1016/0079-6700(92)90003-H. 2.

Glasser, W.G. Classification of lignin according to chemical and molecular structure, in: Glasser, W.G.; Northey, R.A.; Schultz, T.P. (Eds.), Lignin: Historical, Biological, and 20 ACS Paragon Plus Environment

Page 21 of 27 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

403

Materials Perspectives, ACS Symp. Ser., 1999, vol. 742 pp. 216–238, DOI:10.1021/bk-

404

2000-0742.ch009.

405

3. Feldman, D. Lignin and its polyblends – a review, in: Hu, T.Q. (Ed.), Chemical Modification,

406

Properties, and Usage of Lignin, Kluwer Academic/Plenum Publishers, New York, 2002,

407

pp. 81–99.

408

4. Stewart, D.; Lignin as a base material for materials applications: chemistry, application and

409

economics, Ind. Crop. Prod. 2008, 27, 202-207, DOI:10.1016/j.indcrop.2007.07.008.

410

5. Doherty, W.O.S.; Mousavioun, P.; Fellows, C.M. Value-adding to cellulosic ethanol: lignin

411 412

polymers, Ind. Crop. Prod. 2011, 33, 259-276, DOI:10.1016/j.indcrop.2010.10.022. 6. Thakur, V.K.; Thakur, M.K.; Raghavan, P.; Kessler, M.R. Progress in green polymer

413

composites from lignin for multifunctional applications: a review, ACS Sustainable Chem.

414

Eng. 2014, 2, 1072–1092, DOI:10.1021/sc500087z.

415

7. Sen, S.; Patil, S.; Argyropoulos, D.S.; Thermal properties of lignin in copolymers, blends,

416

and composites: a review, Green Chem. 2015, 17, 4862–4887,

417

DOI:10.1039/C5GC01066G.

418

8. Ghosh, I.; Jain, R.K.; Glasser, W.G.; Blends of biodegradable thermoplastics with lignin

419

esters. In: Glasser, W.G., Northey, R.A., Schultz, T.P. (Eds.), Lignin: Historical, Biological,

420

and Materials Perspectives. American Chemical Society, Washington, DC, 2000; pp. 331-

421

350.

422

9. Kadla, J.F.; Kubo, S. Lignin-based polymer blends: analysis of intermolecular interactions

423

in lignin-synthetic polymer blends, Composites Part A. 2004, 35, 395-400,

424

DOI:10.1016/j.compositesa.2003.09.019.

425

10. Kubo, S.; Kadla, J.F. Kraft lignin/poly(ethylene oxide) blends: effect of lignin structure on

426

miscibility and hydrogen bonding. J. Appl. Polym .Sci. 2005, 98,1437-44,

427

DOI:10.1002/app.22245.

21 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

428

11. Kadla, J. F.; Kubo, S.; Venditti, R.A.; Gilbert, R.D.; Compere, A.L.; Griffith, W. Lignin based

429

carbon fibers for composite fiber applications, Carbon, 2002, 40, 2913-2920, DOI:

430

10.1016/S0008-6223(02)00248-8.

431 432

Page 22 of 27

12. Kadla, J. F.; Kubo, S. Miscibility and Hydrogen Bonding in Blends of Poly(ethylene oxide) and Kraft Lignin, Macromolecules, 2003, 36, 7803-7811, DOI:10.1021/ma0348371.

433

13. ) Lewis, N. G.; Lantzy, T. R. In Adhesives from Renewable Resources; Hemingway, R. W.,

434

Conner, A. H.; Branham, S. J., Eds.; ACS Symposium Series 385; American Chemical

435

Society: Washington, DC, 1989; pp 13−26, DOI:10.1021/bk-1989-0385.ch002.

436

14. Zhao Y.; Yan, N., Recent development in forest biomass derived phenol formaldehyde

437

(PF) resol resin for wood adhesives application. J. Biobased mater. Bio. 2014, 8, 465-480,

438

DOI:10.1166/jbmb.2014.1463.

439

15. Kasbe, P. S.; Kumar, N.; Manik, G. A molecular simulation analysis of influence of

440

lignosulphonate addition on properties of modified 2.ethyl hexyl acrylate/methyl

441

methacrylate/ acrylic acid based pressure sensitive adhesive, Int. J. Adhes. Adhes. 2017,

442

78, 45-54, DOI:10.1016/j.ijadhadh.2017.06.014.

443

16. Wang, S.; Shuai, L.; Saha, B.; Vlachos, D. G.; Epps, T.H. III, From Tree to tape: Direct

444

synthesis of pressure sensitive adhesives from depolymerized raw lignocellulosic biomass.

445

ACS Cent. Sci. 2018, 4, 701-708, DOI:10.1021/acscentsci.8b00140.

446 447 448

17. Creton, C. Pressure-Sensitive Adhesives: An Introductory Course. MRS Bulletin. 2003, 28(6), 434-439, DOI:10.1557/mrs2003.124. 18. Gillet, S.; Aguedo, M.; Petitjean, L.; Morais, A. R. C.; da Costa Lopes, A. M.; Łukasik, R.

449

M.; Anastas, P. T. Lignin transformations for high value applications: towards targeted

450

modifications using green chemistry. Green Chem. 2017, 19, 4200-4233,

451

DOI:10.1039/C7GC01479A.

452

19. Aziz S.; Sarkanen, K. Organosolv pulping-a review. Tappi J. 1989, 72, 169–175.

22 ACS Paragon Plus Environment

Page 23 of 27 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

453

ACS Sustainable Chemistry & Engineering

20. Zhang, T.; Gao, J.; Qi, B.; Liu, Y., Performance of concrete made with superplasticizer from

454

modified black liquor and polycarboxylate, Bioresources. 2014, 9 (4), 7352-7362,

455

DOI:10.15376/biores.9.4.7352-7362.

456

21. Gupta, C.; Nadelman, E.; Washburn, N. R.; Kurtis, K.E. Lignopolymer superplasticizers for

457

low-CO2 cements. ACS Sustainable Chem. Eng. 2017, 5, 4041-4049,

458

DOI:10.1021/acssuschemeng.7b00021.

459

22. MA, S.; Ding, Q.; Zhou F.; Zhu, H. Synthesis and characterization of lignin grafting

460

modification-based aliphatic superplasticizer, J. of Wuhan University of Technology-Mater.

461

Sci. Ed. 2018,33, 661, DOI:10.1007/s11595-018-1875-z.

462 463 464

23. Aubrey, D,W.; Sherriff, M.; Peel adhesion and viscoelasticity of rubber–resin blends. J. Appl. Polym. Sci. 1980, 18, 2597-2608, DOI: 10.1002/pol.1980.170180818. 24. 2. Sherriff, M.; Knibbs, R.W.; Langley, P.G. Mechanism for the action of tackifying resins in

465

pressure-sensitive adhesives. J. Appl. Polym. Sci. 1973,17, 3423-3438,

466

DOI:10.1002/app.1973.070171114

467 468 469

25. Hata,T.; Tsukatani, T.; Mizumachi, H. Relationship between holding power and sliding friction coefficient of pressure sensitive adhesives. J. Adhes. Soc. Jpn.1994, 30, 307-312. 26. Over, L. C.; Grau, E.; Grelier, S.; Meier, M. A. R.; Cramail, H. Synthesis and

470

Characterization of Epoxy Thermosetting Polymers from Glycidylated Organosolv Lignin

471

and Bisphenol A. Macromol. Chem. Phys. 2017, 218,1600411,

472

DOI:10.1002/macp.201600411.

473

27. Le Floch A.; Jourdes M.; Teissedre P. L.; Polysaccharides and lignin from oak wood used

474

in cooperage: Composition, interest, assays: a review: Carbohydr. Res. 2015, 417, 94-102,

475

DOI: 10.1016/j.carres.2015.07.003.

476 477 478

28. ASTM, D 3121-94, standard test method for tack of pressure-sensitive adhesives by rolling ball. 29. http://www.packtest.com/wp-content/uploads/2017/01/Rolling-Ball-Tack-Tester.pdf. 23 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

479

30. P. S. T. Council, “PSTC 101 test method: peel adhesion of pressure sensitive tape,” Test

480

Methods for Pressure Sensitive Tapes, 14th Edition. Northbrook, IL: Pressure Sensitive

481

Tape Council, 2004, 101,1–101.

482

31. C. S.-S. Sheet, “ASTM A 240/A 240M or ASTM A 666,” Type , 304, 25.

483

32. P. S. T. Council: PSTC-107 test method: International Standard for Shear Adhesion of

484 485

Pressure Sensitive Tape. 33. Zhao J.; Ediger, M. D.; Sun, Y.; Yu, L. Two DSC Glass Transitions in Miscible Blends of

486

Polyisoprene/Poly(4-tert-butylstyrene), Macromolecules, 2009, 42, 6777–6783,

487

DOI:10.1021/ma9008516.

488 489 490 491 492 493 494

34. Coleman, M. M.; Graf, J. F.; Painter, P. C. Specific Interactions and the Miscibility of Polymer Blends. Technomic Publishing, Lancaster, PA, USA,1991. 35. Fox, T. G. Influence of diluent and of copolymer composition on the glass temperature of a polymer system. Bull. Am. Phys. Soc. 1956,1, 123. 36. Gordon, M.; Taylor, J. Ideal copolymers and the second-order transitions of synthetic rubbers. i. Non-crystalline polymers. J. Appl. Chem.1952, 2, 493-500. 37. Kwei, T. K.; Pearce, E. M.; Pennacchia, J. R.; Charton, M. Correlation between the Glass

495

Transition Temperatures of Polymer Mixtures and Intermolecular Force Parameters

496

Macromolecules 1987, 20, 1174-1176,DOI: 10.1021/ma00171a055.

497

38. Weng, L.; Vijayaraghavan, R.; MacFarlane, D.R.; Elliott, G. D. Application of the Kwei

498

Equation to model the Tg Behavior of Binary Blends of Sugars and Salts, Cryobiology.

499

2014, 68, 155-158, DOI:10.1016/j.cryobiol.2013.12.005.

500

Page 24 of 27

39. Kuo, S. W.; Chang, F. C. Miscibility and hydrogen bonding in blends of poly (vinylphenol-

501

co-methyl methacrylate) with poly (ethylene oxide), Macromolecules, 2001, 34, 4089-4097,

502

DOI:10.1021/ma010047k.

24 ACS Paragon Plus Environment

Page 25 of 27 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

503

ACS Sustainable Chemistry & Engineering

40. Kwei, T. K. The effect of hydrogen bonding on the glass transition temperatures of polymer

504

mixtures. J. Polym. Sci, Polym. Lett. Ed. 1984, 22, 307-313,

505

DOI:10.1002/pol.1984.130220603.

506

41. Zhao, X.; Zhang, Y.; Wei, L.; Hu, H.; Huang, Z.; Yang, M.; Huang, A.; Wu, J.; Feng, Z.

507

Esterification mechanism of lignin with different catalysts based on lignin model compounds

508

by mechanical activation-assisted solid-phase synthesis. RSC Adv., 2017, 7, 52382-52390,

509

DOI: 10.1039/C7RA10482K.

510 511 512 513 514 515

42. R.M. Ottenbrite, L.A. Utracki, S. Inoue (Eds.), Current topics in polymer science, rheology and polymer processing/multiphase systems, vol. II, Carl Hanser, Munich (1987), pp. 7-59. 43. Yang, H. W. H.; Chang, E. P. The Role of Viscoelastic Properties in the Design of Pressure-sensitive Adhesives. Trends Polym. Sci. 1997, 5, 380-384. 44. Dahlquist, C.A. Pressure-Sensitive Adhesives. In Treatise on Adhesion and Adhesives; Dekker, New York, 1969, 2, 219-260.

516

45. Deplace, F.; Carelli, C.; Mariot, S.; Retsos, H.; Chateauminois A.; Ouzineb,K.; Creton, C.

517

Fine Tuning the Adhesive Properties of a Soft Nanostructured Adhesive with Rheological

518

Measurements. J. Adhes. 2009, 85, 18-54, DOI:10.1080/00218460902727381.

519

46. Callies, X.; Herscher, O.; Fonteneau, C.; Robert, A.; Pensec, S.; Bouteiller, L.;

520

Ducouret,G.; Creton, C. Combined effect of chain extension and supramolecular

521

interactions on rheological and adhesive properties of acrylic pressure-sensitive adhesives,

522

ACS Appl. Mater. Interfaces 2016, 8, 33307−33315. DOI:10.1021/acsami.6b11045.

523

47. Zhang, K.; Fahs, G. B.; Margaretta, E.; Hudson, A. G.; Moore, R. B.; Long, T. E. Acetyl-

524

Protected Cytosine and Guanine Containing Acrylics as Supramolecular Adhesives . J.

525

Adhes. 2019, 95, 146–167, DOI:10.1080/00218464.2017.1419430.

526 527

48. Chang, E.P. Viscoelastic windows of pressure-sensitive adhesives. J. Adhes. 1991, 34,189-200, DOI:10.1080/00218469108026513.

25 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

528

Page 26 of 27

49. Lee, S.; Lee, K.; Kim, Y.W.; Shin, J. Preparation and Characterization of a Renewable

529

Pressure-Sensitive Adhesive System Derived from ε-Decalactone, l-Lactide, Epoxidized

530

Soybean Oil, and Rosin Ester, ACS Sustainable Chem. Eng. 2015, 3, 2309-2320,

531

DOI:10.1021/acssuschemeng.5b00580.

532

50. Gallagher, J. J.; Hillmyer, M. A.; Reinecke, T. M. Acrylic Triblock Copolymers Incorporating

533

Isosorbide for Pressure Sensitive Adhesives. ACS Sustainable Chem. Eng. 2016, 4, 3379-

534

3387, DOI:10.1021/acssuschemeng.6b00455.

535

51. Callies, X.; Fonteneau, C.; Pensec, S.; Bouteiller, L.; Ducouret,G.; Creton, C. Adhesion

536

and non-linear rheology of adhesives with supramolecular crosslinking points, : Soft Matter,

537

2016, 12, 7174-7185, DOI: 10.1039/c6sm01154c.

538

52. Callies, X.; Vechamber, C.; Fonteneau, C.; Pensec, S.; Chenal, J. M.; Chazeau, L.;

539

Bouteiller, L.; Ducouret,G.; Creton, C. Linear rheology of supramolecular polymers center-

540

functionalized with strong stickers, Macromolecules 2015, 48, 7320−7326,

541

DOI:10.1021/acs.macromol.5b01583.

542

53. Cashion, M. P.; Park, P.; Long, T. E. Influence of hydrogen bonding on the adhesive

543

properties of photo-curable acrylics, J. Adhes, 2009, 85:1–17,

544

DOI:10.1080/00218460902727332.

545

54. Kajtna, J.; Alič, B.; Krajnc, M.; Šebenik, U. Influence of hydrogen bond on rheological

546

properties of solventless UV crosslinkable pressure sensitive acrylic adhesive prepolymers,

547

Int. J. Adhes & Adhes, 2014, 49, 103–108, DOI:10.1016/j.ijadhadh.2013.12.016.

548 549 550 551 552 553 26 ACS Paragon Plus Environment

Page 27 of 27 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

554

ACS Sustainable Chemistry & Engineering

Graphical Abstract: For Table of Contents Use Only

555 556 557

Brief Synopsis

558

The potential use of organosolv lignin in the base formulation of pressure sensitive adhesive was

559

assessed by its adhesive performances.

560

27 ACS Paragon Plus Environment