Cross-linked and shapeable porous 3D substrates from freeze-linked

32 mins ago - ... have been prepared using a freeze-linking procedure which avoids common post activation of crosslinking reactions and freeze-drying...
0 downloads 0 Views 5MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Cross-linked and shapeable porous 3D substrates from freeze-linked cellulose nanofibrils Johan Erlandsson, Hugo Françon, Andrew Marais, Hjalmar Granberg, and Lars Wågberg Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.8b01412 • Publication Date (Web): 05 Nov 2018 Downloaded from http://pubs.acs.org on November 6, 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 41 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

Biomacromolecules

1

Cross-linked and shapeable porous 3D substrates

2

from freeze-linked cellulose nanofibrils

3

Johan Erlandsson a*, Hugo Françon a, Andrew Marais a, Hjalmar Granberg b, Lars Wågberg a,c*

4

a. Division of Fibre Technology at the Department of Fibre and Polymer Technology, School of

5

Engineering Sciences in Chemistry, Biotechnology and Health KTH Royal Institute of

6

Technology, Teknikringen 56-58, SE-100 44, Stockholm, Sweden

7

b. Papermaking & Packaging, RISE Bioeconomy, Box 5604, SE-114 86, Stockholm, Sweden

8

c. Wallenberg Wood Science Center, KTH Royal Institute of Technology, Teknikringen56, SE-

9

100 44, 8 Stockholm, Sweden

10

11 12 13 14 15 16

17

ACS Paragon Plus Environment

1

Biomacromolecules 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

18

Page 2 of 41

Abstract

19

Chemically crosslinked highly porous nanocellulose aerogels with complex shapes have been

20

prepared using a freeze-linking procedure which avoids common post activation of crosslinking

21

reactions and freeze-drying. The aerogel shapes ranged from simple geometrical three

22

dimensional bodies to swirls and solenoids. This was achieved by moulding or extruding a

23

periodate oxidized cellulose nanofibril (CNF) dispersion prior to chemical crosslinking in a

24

regular freezer, or by reshaping an already prepared aerogel by plasticizing the structure in water

25

followed by reforming and locking the aerogel into its new shape. The new shapes were most

26

probably retained by new crosslinks formed between CNFs brought into contact by the

27

deformation during reshaping. This self-healing ability to form new bonds after plasticization

28

and redrying also contributed to the mechanical resilience of the aerogels, allowing them to be

29

cyclically deformed in the dry state, reswollen with water and redried with good retention of

30

mechanical integrity. Furthermore, by exploiting the shapeability and available inner structure of

31

the aerogels, a solenoid shaped aerogel with all surfaces coated with a thin film of conducting

32

polypyrrole was able to produce a magnetic field inside the solenoid demonstrating

33

electromagnetic properties. Also, by biomimicking the porous interior and stiff exterior of the

34

beak of a toucan bird, an aerogel functionalization was created by applying a 300 µm thick stiff

35

wax coating on its moulded external surfaces. This composite material displayed a ten times

36

higher elastic modulus compared to the plain aerogel without drastically increasing in density.

37

These examples show that it is possible to combine advanced shaping with functionalisation of

38

both the inner structure and the surface of the aerogels, radically extending the possible use of

39

aerogels.

40

ACS Paragon Plus Environment

2

Page 3 of 41 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

41

Biomacromolecules

Introduction

42

Cellulose nanofibrils (CNFs) and cellulose nanocrystals (CNCs) are renewable, bio-derived

43

nanomaterials that can easily be converted into several types of macroscopic advanced materials

44

such as gas barrier films,1-3 high strength filaments,4-5 low density foams

45

Nanocellulose aerogels are commonly prepared by freeze-drying (FD) or critical point drying

46

(CPD), and properties such as density, specific surface area, pore structure and mechanical

47

properties can be tuned by altering the preparation procedure.8,

48

content of native hydroxyl groups and carboxyl groups introduced during pre-treatments, to

49

facilitate the fibril liberation,13-14 makes CNF aerogels post-functionalizable both chemically and

50

physically to produce new functional materials.6, 15 However, as cellulose is sensitive to water,

51

nanocellulose aerogels tend to easily disintegrate when soaked in water, limiting both the

52

possibility of using subsequent water-based functionalization methods and final applications.16-17

53

To address this issue several methods for chemical crosslinking of nanocellulose aerogels

54

through e.g. ester bonds by using tetrafunctional carboxylic acids or by commercial crosslinking

55

agents such as Kymene, have been developed, giving rise to water-stable aerogels.15, 17-18 The

56

wet-stability also allows functionalization methods such as the water-based Layer-by-Layer

57

(LbL) method, used to prepare aerogels with properties ranging from flame-retardancy,

58

antibacterial activity, cell compatibility and energy storage.18-21

11-12

6-7

and aerogels.8-10

Furthermore, the high

59

Despite the versatility of nanocellulose aerogels as templates for functionalization and the

60

broad range of applications they offer, upscaling remains a challenge because of the use of FD

61

and CPD. Moreover, their use in shape demanding applications is limited because of their lack of

62

shapeability, mainly due to their inherent brittleness and friability. The achievable shapes are

63

generally limited to relatively simple geometries due to the use of moulds that need to be

ACS Paragon Plus Environment

3

Biomacromolecules 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 41

64

removed without damaging the aerogel. Hence, complex shapes such as spirals and loops cannot

65

be achieved. In a recent study a procedure to fabricate and shape wet-stable CNF aerogels

66

without any freeze-drying, using a simple freeze-thawing and solvent exchange procedure was

67

reported and opened up possibilities of an easier and more versatile production.22 The produced

68

aerogels could be soaked in water and compressed to 70% without losing their ability to recover

69

their original shape and size. The wet stability was achieved by crosslinking through hemiacetal

70

bonds between aldehydes introduced to the CNFs by sodium metaperiodate oxidation and

71

hydroxyls on neighbouring CNFs.23

72

Following up this work, we herein describe three methods how these freeze-linked aerogels

73

can be prepared into complex shapes using different shaping procedures. We also show that the

74

structural properties (density, porosity, pore size) of the aerogels can be tuned by varying the

75

CNF concentration. Furthermore, we provide new properties to the aerogels through

76

functionalization of both the inner structure and the outer surface of the aerogels. The

77

functionalization is done with both inactive and active components such as natural waxes and

78

conducting polymers. Inspired by the toucan bird, we use a biomimicking approach to prepare

79

stiff and lightweight core-shell composites resembling the toucan beak. Finally by

80

polymerisation of pyrrole within the porous structure of a solenoid-shaped freeze-linked aerogel,

81

we show that it is possible to create a device that can generate a magnetic field.

82

Materials and Methods

83

Materials

84

Carboxymethylated CNFs were produced according to a previously reported method

14

and

85

were provided by RISE Bioeconomy AB, Stockholm, Sweden, as a 20 g/L gel. Sodium

86

metaperiodate, sodium chloride and hydroxylamine hydrochloride, carnauba wax, pyrrole,

ACS Paragon Plus Environment

4

Page 5 of 41 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

Biomacromolecules

87

anhydrous iron(III) chloride, para-toluenesulfonic acid were all purchased from Sigma Aldrich.

88

All chemicals were used without further purification unless otherwise stated.

89

Methods

90

Preparation of crosslinked CNF aerogels by freeze-induced crosslinking

91

CNF aerogels were prepared by mixing sodium periodate and CNF of varying concentrations

92

in a 0.7:1 dry weight ratio. The oxidation reaction was carried out for 1 hour after which the

93

CNF-periodate mixture was transferred to a pre-selected mould and then placed in a freezer (-18

94

°C) over night. The frozen samples were then thawed at room temperature, solvent exchanged to

95

acetone and dried under ambient conditions.

96

Total charge determination

97

Conductometric titration was used to determine the total carboxyl content of the material and

98

was performed with a Titrino 702 SM (Metrohm AG, Herisau, Switzerland) according to a

99

previously described method 24. Prior to the measurements, the CNF aerogels were washed with

100

water to remove any residual salt remaining after the oxidation reaction. The aerogels were

101

washed until the conductivity of the washing water became lower than 5 µS/cm. The carboxyl

102

groups present in the materials were subsequently converted into their proton form by

103

equilibration in water at pH 2 for 30 min after which they were again washed with water until the

104

conductivity of the washing water became lower than 5 µS/cm. 0.1 g of CNF aerogel material

105

(dry weight), was added to a total volume of 500 mL of water containing 0.1 mM HCl and 2 mM

106

NaCl. The solution was purged from oxygen and carbon dioxide with nitrogen gas for 15 min

107

prior to the titration. The titration was performed with 0.1 M NaOH and the amount of titrant

ACS Paragon Plus Environment

5

Biomacromolecules 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 41

108

used for calculation of the total charge was determined from the plateau region for the

109

conductivity in the titration curve 24.

110 111

Aldehyde content determination

112

The amount of aldehydes introduced into the CNFs by the periodate oxidation was determined

113

by titration with sodium hydroxide after reaction with hydroxylamine hydrochloride, which

114

reacts with the aldehydes forming oximes while releasing a stoichiometric amount of protons 25.

115

The oxidized material was added to 25 mL of water containing 10 mM NaCl to ensure that the

116

pH on the surface of the fibrils was the same as that in the bulk, and the pH was adjusted to 4.

117

The sample mixture was then mixed with 25 mL of 0.25 M hydroxylamine hydrochloride

118

solution (also containing 10 mM NaCl) adjusted to pH 4, and the mixture was allowed to react

119

for 2 h prior to a back titration at pH 4 using 0.1 M NaOH. The amount of aldehydes in the

120

sample was calculated from the number of moles of NaOH needed to reach pH 4 and the results

121

were normalized with respect to the dry weight of the CNF sample.

122

Morphology

123

The microstructure and morphology of the aerogels produced was studied using an S-4800

124

field emission scanning electron microscope (SEM) (Hitachi, Tokyo, Japan). Aerogels samples

125

were carefully cut and glued onto a sample holder using conducting carbon tape. Prior to

126

imaging the samples were coated with Pt/Pd in a Cressington 208 HR sputter coater (Cressington

127

Scientific Intruments, Watford, UK) for 40 s to limit specimen charging during imaging in the

128

SEM. The pore size was estimated from SEM images as an average of the pore diameter

129

measured in two perpendicular directions.

130

ACS Paragon Plus Environment

6

Page 7 of 41 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

131

Biomacromolecules

Mechanical properties

132

The aerogels were tested at 50% RH and 23 °C in the dry state using an Instron 5566 universal

133

testing machine (Norwood, MA, USA) equipped with a 500 N load cell. The samples consisted

134

of cubic aerogel samples of approximatively 20 mm side (the exact dimensions were determined

135

with a caliper before testing). The rate of compressive strain was 10 %/min, and the samples

136

were compressed until they reached 80% compression. Unless stated otherwise, six samples were

137

tested for each aerogel type.

138

Wax coating

139

Cubic aerogels (17.5 g/L, 2x2x2 cm3) were coated with carnauba wax on four adjoining sides

140

by spin coating, thus leaving two opposite uncoated faces. The wax was melted and set to a

141

temperature of 105 °C and subsequently added dropwise, a total of five drops per side, to the

142

spinning aerogel (3600 rpm for 20 s). Excess wax extending off the sides of the aerogel was

143

removed using a heated spatula. The coated aerogels were tested in compression with the coated

144

sides in the compression direction.

145

Polypyrrole functionalization

146

The polypyrrole (PPy) functionalization of the porous aerogel structure was performed by

147

soaking the aerogel in a water solution containing pyrrole, p-Toluenesulfonic acid and

148

iron(III)chloride and allowing the polymerization to continue for 24h. The initial concentration

149

of pyrrole was 0.033 M and the ratio of p-Toluenesulfonic acid/pyrrole was 0.3 and,

150

iron(III)chloride/pyrrole 2.3 respectively. Following the polymerization, the aerogel was washed

151

extensively with water and subsequently solvent exchanged to acetone and ambiently dried. Any

152

electrical connections to the aerogel were managed by gluing copper wires to the aerogel using

ACS Paragon Plus Environment

7

Biomacromolecules 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 41

153

silver paint. The electrical properties of the aerogel were evaluated using a Keithley 2400 series

154

Source-Meter (Keithley Instruments Inc., OH, USA).

155

Magnetic properties

156

The magnetic properties were measured using an Adafruit HMC5883L tripe axle

157

magnetometer connected to an Arduino Uno microcontroller with the Adafruit sensor library.

158

The magnetometer was inserted into the middle of a solenoid shaped aerogel functionalized with

159

PPy with one axis parallel to the solenoid. The magnetic field was subsequently recorded

160

continuously while current passing through the spiral was alternatively turned on and off.

161

162

Results & Discussion

163

Principles of CNF aerogel preparation

164

The freeze-linked CNF aerogels all shared one common production procedure shown in a flow

165

chart in Figure 1a. The only difference is found in the shaping where one of three different

166

methods, moulding, extruding and reshaping, were used to achieve the final shape of the

167

material. The moulding and extruding methods (Fig 1b) shaped the CNF dispersion into the

168

desired shape prior to crosslinking the structure by freezing. The third shaping method involved

169

reshaping and locking an already prepared aerogel, i.e. an already shaped aerogel (through

170

moulding or extrusion) into a new shape, as exemplified in Fig 1c.

171

The freeze-linking technique relies, similarly to freeze-casting,26-28 on the exclusion of the

172

CNFs and the solutes from the growing ice crystals and their packing in the thin lamellae

173

between the crystals during freezing, displayed schematically in Fig 1d. It is in these lamellae in

174

the frozen state that the CNFs crosslink through hemiacetal linkages and produce a wet-stable

ACS Paragon Plus Environment

8

Page 9 of 41 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

Biomacromolecules

175

structure.29 The creation of the 3D structure by ice-templating a colloid is dependent on the

176

nucleation and growth of the ice crystals which in turn are dependent on for example the

177

ice/solid interactions, the degree of super cooling

178

both the amount of periodate used during the oxidation and the shape of the extruded body and

179

mould affect the freezing behaviour and therefore yield different pore structures. The wide range

180

of herein reported aerogel shapes therefore give rise to pore structures which are strongly related

181

to their initial shape, CNF concentration and degree of oxidation.

30

and the ionic strength and ion type.31 Thus

182 183

Figure 1 (a) Schematic representation of the preparation of the aerogels, (b) shaping by moulding

184

or extrusion (c) post production reshaping of aerogels, (d) schematic view of the ice crystals

185

excluding the CNFs from the ice-phase and tightly packing them together in the lamellae.

ACS Paragon Plus Environment

9

Biomacromolecules 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 41

186

Aerogels were prepared from CNF dispersions at concentrations ranging from 7.5 g/L to

187

20 g/L, which are well above the overlapping concentration (0.4-0.7 g/L) 32 of the CNFs used in

188

this study, also after the periodate oxidation. Figure 2a shows a 5 cm diameter spherical aerogel

189

shaped by moulding of a 20 g/L CNF dispersion together with its porous microstructure which

190

consists of 2D sheets of compressed CNFs (Fig 2b), also seen at higher magnification in

191

Figure 2c. Below 7.5 g/L, the solvent exchanged samples collapsed during ambient drying. Since

192

a decreased CNF concentration resulted in fewer CNFs packing together in the lamellae during

193

freezing, the collapse was presumably due to insufficient mechanical strength of the formed

194

structure. It has also been reported that freezing of CNFs at 5 g/L and lower concentrations

195

results in a weak fibrillar structure that collapses (even if freeze-drying was used) rather than in a

196

stronger structure of interconnected 2D sheets.33 When soaking aerogels with the fibrillar and

197

sheet like structures in water the aerogels with the fibrillar structure collapsed due to capillary

198

forces while the aerogels with the 2D sheets remained intact34. Examination of a 0.1 g/L CNF

199

dispersion that was oxidized and freeze-dried showed that the oxidized CNFs also formed a

200

fibrillar structure, as shown in Figure S1. The collapse of the aerogels with a concentration

201

below 7.5 g/L is therefore presumably due to a combination of both low mechanical strength of

202

the formed CNF structure and possibly also to a different 3D structure resulting from the low

203

CNF concentration. The effect of the CNF dispersion concentration on how the CNF is

204

templated in the lamellae between the crystals clearly demonstrates the wide range of pore

205

structures possible to produce by ice templating . The formation of the 2D-sheets was evidently

206

required to obtain a structure with sufficient strength to withstand the ambient drying step

207

following solvent exchange.

ACS Paragon Plus Environment

10

Page 11 of 41 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

Biomacromolecules

208

Figure 2d shows the evolution of the aerogel density as a function of the initial CNF

209

concentration. A natural consequence of using a range of CNF concentrations was that the

210

density of the aerogels increased, from 12 kg/m3 to 34 kg/m3, (between 98-99% porosity) as the

211

concentration increased from 7.5 g/L to 20 g/L. Although the aerogels were dried at room

212

temperature and subjected to capillary forces, the average volumetric retention of the aerogels

213

was above 75% regardless of the initial CNF concentration. This value is comparable to

214

volumetric retentions reported for CNF based aerogels produced from similar CNF

215

concentrations and dried by freeze-drying 12. All samples displayed wet-integrity when soaked in

216

water and were able to be redried after a second solvent exchange, In Figure S2 two aerogels

217

(7.5 g/L and 17.5 g/L) that have been completely soaked in water without disintegrating are

218

displayed. This suggested that the crosslinking through hemiacetals between the CNFs in the

219

lamellae was effective also at the lowest CNF concentration. This is in line with the aldehyde

220

content of all samples, being between 1.0 mmol/g and 1.1 mmol/g, which exceeded the

221

0.6 mmol/g, previously reported as the threshold value needed for effective crosslinking 29. The

222

total charge (carboxylic acid content) of the final material was on average 350 µeq/g,

223

independent of CNF concentration, which is significantly lower than the initial 600 µeq/g. This

224

agrees with previous studies on aerogels with similar aldehyde contents produced using the same

225

freeze-linking procedure. The loss of charged groups in this case was attributed to dissolution of

226

the charged and highly oxidized cellulose chains on the surface of the fibrils 22, 35-36.

ACS Paragon Plus Environment

11

Biomacromolecules 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 41

227 228

Figure 2 (a) Photograph of a spherical cross-linked CNF aerogel made from 20 g/L CNF gel

229

(scale bar = 1 cm), (b) SEM-image of the cross-section of the spherical aerogel (scale bar =

230

500 µm), (c) SEM image showing the pore walls (scale bar = 10 µm), (d) The aerogel density as

231

a function of the initial CNF concentration.

232

Shaping the aerogels through moulding limited the geometries to simple geometrical bodies

233

such as spheres (Fig 2a), however, by using a CNF dispersion with a concentration of 17.5 g/L or

234

above shaping of the CNF dispersion by extrusion was possible. The viscosity of the dispersions

235

above 17.5 g/L was sufficient to allow the dispersion to retain the nozzle shape also after

236

extrusion. Extrusion of the CNF gel through a star-shaped nozzle yielded shapes such as the one

237

displayed in Fig 3a. The possibility to extrude the dispersions into a specific shape opens up the

ACS Paragon Plus Environment

12

Page 13 of 41 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

Biomacromolecules

238

possibility of 3D-printing advanced shapes and subsequently crosslink the printed structure in a

239

consecutive freezing step.

240

The shape of the aerogels could also be altered post-preparation by soaking the aerogel in

241

water to plasticize the structure, reshaping it and subsequently locking the structure into a new

242

shape. The new shape can be locked by submitting the strained structure to either a new solvent

243

exchange step with subsequent drying or by another freezing, solvent exchange and drying

244

sequence. To illustrate this reshapability property, a solenoid shaped aerogel was prepared, see

245

Figure 3b. The solenoid spring was obtained by soaking the aerogel in water and wrapping it

246

around a tube before solvent-exchange and drying. The ability to be re-soaked, solvent

247

exchanged and dried indicates that the structure is largely undamaged by the physical treatment

248

involved in reshaping, and that the strength of the pore walls remained also after being reshaped.

249

The reshapeability was further tested by soaking an aerogel in water and inserting the softened

250

structure in a different mould, freezing, solvent exchange and drying. Figure 3c shows a

251

cylindrical aerogel that was reshaped into a prism following this method.

ACS Paragon Plus Environment

13

Biomacromolecules 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 41

252 253

Figure 3 Photographs of (a) an extruded swirl aerogel, (b) a solenoid- shaped aerogel prepared by

254

reshaping a long cylindrical aerogel , (d) a cylindrical aerogel (left) reshaped into a prism (right).

255

The scale bars are 1 cm.

256

Given that the fibrils in the aerogel are crosslinked and therefore fixed in space with respect to

257

each other and that the structure is relatively inelastic in tension, a change of shape requires a

258

change in volume. Hence, a change in density naturally follows due to the densification of the

259

cylinder when it was reshaped into a prism. The final density of the prism was 77 kg/m3 and the

260

initial of the cylinder 34 kg/m3. Note that the density increase was due to the deformation and

261

densification of the outer parts of the cylinder and not due to an overall densification. The

262

densified parts of the aerogel were visible as skin on the outer parts of the aerogel. This

263

demonstrates the possibility of creating aerogels with density gradients. As the new structure

264

contains cell walls that have been deformed in a plasticized state and subsequently pressed

265

together by both the mechanical force induced by the decreased volume of the mould and by the

266

ice forming in the pores, the aerogel is crosslinked into the new shape. This was evident from

ACS Paragon Plus Environment

14

Page 15 of 41 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

Biomacromolecules

267

soaking the aerogel in water where it maintained its prism structure and did not revert to its

268

original cylindrical shape even after soaking in water for several hours.

269

Microstructure and morphology of the moulded freeze-linked aerogels

270

All moulded, extruded or reshaped aerogels, displayed an open pore structure with

271

interconnected 2D sheet-like pore walls comprising oxidized CNFs tightly packed into a CNF

272

nano sheet structure3, irrespective of CNF concentration from 7.5 g/L and above. Overall, the

273

pores were not found to be aligned, of a specific shape or oriented. Instead all were of random

274

size and orientation (Fig 4). This suggested that the nucleation of ice crystals was randomly

275

distributed with no preferred growth direction in the dispersion during the freezing, unlike the

276

freeze casting method27 where the crystals are continuously growing parallel to a temperature

277

gradient.

278

In the aerogels prepared from the lowest CNF concentration (7.5 g/L) the pores were

279

significantly larger compared to the pores found in the aerogels made from higher CNF

280

concentrations, see Fig 4a. Presumably, at lower solid contents the CNFs are not blocking each

281

other and are easily excluded from the propagating ice front as the ice crystals grow larger.30

282

Smaller pores of similar size were observed in the aerogels prepared at higher concentrations

283

(Fig 4b and c).

284 285

Figure 4. SEM images of cross sections of moulded (cubic) aerogels prepared from (a) 7.5 g/L,

286

(b) 17.5 g/L and (c) 20 g/L. Scale bars represent 1 mm.

ACS Paragon Plus Environment

15

Biomacromolecules 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 41

287

At higher CNF concentrations the excluded CNFs accumulate at the interface of the growing

288

ice front and start forming pore walls earlier which hinders the propagation of the ice front and

289

hence creates smaller pores. In addition a higher CNF content also allows for an increased

290

number of sites where heterogeneous nucleation of ice crystals can occur, which also would

291

decrease the pore size. As discussed earlier, several parameters affects the formation of the pore

292

structure in the aerogel and as the ionic strength varied (since it is based on the CNF solid

293

content) and the oxidation reaction altered the CNF/CNF and CNF/water interactions29, it is

294

difficult to isolate and quantify the contribution of each of those parameters in the formation of

295

the pore structure.

296 297 298

Compressive properties of freeze-linked aerogels

299

The mechanical properties of the freeze-linked aerogels were evaluated in uniaxial

300

compression of cubic aerogel samples. All samples displayed a typical aerogel behaviour with

301

three different regions in the compression curve: an elastic deformation at 60% where opposite

304

cell walls come into contact and the stress increases due to the densification of the overall

305

structure. Figure 5a displays typical stress-strain curves in uniaxial compression.

ACS Paragon Plus Environment

16

Page 17 of 41 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

Biomacromolecules

306 307

Figure 5. (a) Compressive stress strain curves for the freeze-linked aerogels prepared from

308

different initial CNF concentrations. The inset shows the initial part of the compression test and

309

(b) the cell wall properties calculated using the Ashby and Gibson scaling laws.

310

In general, as the CNF concentration increased, the compressive strength of the aerogel is

311

increased. This is similar to the mechanical behaviour reported for CNF aerogels previously11

312

and a summary of the mechanical properties is presented in Table 1. However, aerogels prepared

313

from 17.5 g/L dispersions exhibited both higher elastic modulus and yield stress compared to the

314

20 g/L aerogels. This observation of the significantly higher modulus, ~1000 kPa (17.5 g/L),

315

which is twice that at 15 g/L and 20 g/L (~500 kPa), suggests that it is not only the concentration

316

of CNFs that is contributing to the final mechanical characteristics of the aerogel. This somewhat

317

unexpected result was further tested by comparing the results from cubic aerogels with

318

cylindrical aerogels where the same sharp increase in mechanical properties was found for

319

aerogels prepared from a 17.5 g/L CNF dispersion. This excluded the possibility of it being a

320

result of only the geometry of the sample and thus the pore structure governing the mechanical

321

properties. In favor of another mechanism rather than the concentration and geometry is that all

322

samples had the same amount of aldehydes which should allow for equal number of crosslinks

ACS Paragon Plus Environment

17

Biomacromolecules 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 41

323

between the CNFs and thus should the property scale only with the number of CNFs. To test this

324

hypothesis the Gibson and Ashby

325

structures were employed.

326

Table 1 Bulk mechanical properties of CNF aerogels prepared from different CNF

327

concentrations, presented as the average value and standard deviation of six samples.

CNF concentration Elastic modulus

37

scaling laws for the mechanical properties of open cell

7.5 g/L

10 g/L

15 g/L

17.5 g/L

20 g/L

101 ± 39

240 ± 83

495 ± 81

1033 ± 240

501 ± 86

4.5 ± 1.6

15.9 ± 3.5

28.1 ± 4.4

38.6 ± 7.7

30.5 ± 5.9

(kPa) Yield strength (kPa) 328 329

The Gibson and Ashby scaling laws suggest that the yield stress and elastic modulus of the cell

330

walls of open cellular materials can be related simply to the relative density of the bulk material (

331

𝜌 ∗ ) and the solids (𝜌𝑠) according to the following relationships:

332

3/2

𝜎∗ 𝜌∗ = 0.3 𝜎𝑠 𝜌𝑠

( )

333

334

𝐸∗ 𝜌∗ = 𝐸𝑠 𝜌𝑠

2

( )

335

Where 𝜎 ∗ , 𝐸 ∗ are the yield stress and elastic moduli of the bulk material and 𝜎𝑠 and 𝐸𝑠 are the

336

yield stress and elastic moduli of the cell walls. The cell wall density (𝜌𝑠) is estimated to be

337

1500 kg/m3 for the CNF walls. This assumption overestimates the cell wall property as the cell

ACS Paragon Plus Environment

18

Page 19 of 41 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

Biomacromolecules

338

wall is not perfectly packed cellulose I but given the resemblance of the cell wall to a CNF-nano

339

sheet which has densities close to 1500 kg/m3 3 this assumption is reasonable. Furthermore, the

340

changes in morphology and complexity of the ice-templated structure also results in that only

341

qualitative estimates of 𝐸𝑠 and 𝜎𝑠were calculated by using the measured bulk densities and

342

mechanical properties. The results are presented in Figure 5b.

343 344

The estimated cell wall modulus reached a maximum value of approximately 4 GPa for the

345

aerogels prepared at 17.5 g/L while the aerogels from 7.5 g/L, 10 g/L, 15 g/L are all resulted in

346

approximately around 2 GPa and the aerogel prepared from 20 g/L had a cell wall modulus of 1

347

GPa. The trend visible in the yield stress of the cell walls also display a maximum at 17.5 g/L

348

and a significant drop for the aerogels prepared at 20 g/L CNF concentration. The almost

349

constant modulus observed below 17.5 g/L is reasonable, given that the CNFs have the same

350

aldehyde content and hence the same properties on an individual CNF level and therefore the

351

global performance only scales with the number of CNFs, i.e. the density as suggested by the

352

scaling laws. Given that the mechanical performance was unaffected by the geometry/pore

353

structure, the sudden increase and subsequent drop observed for the 17.5 g/L and 20 g/L aerogels

354

is presumably due to the CNFs interacting differently at higher concentrations during the

355

freezing process. A possible explanation is that at the lower concentration, the packing and

356

interaction between the CNFs are not limited by interactions with other CNFs when they are

357

excluded from the ice, and can therefore pack together and crosslink to a large extent and the

358

property therefore scales with the amount of CNFs. At 20 g/L the movements of the CNFs

359

become limited as they are excluded from the ice at an earlier stage during freezing which might

360

inhibit their packing and interaction. The 17.5 g/L CNFs dispersions instead presents conditions

ACS Paragon Plus Environment

19

Biomacromolecules 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 41

361

such as ionic strength, CNF concentration and freezing rate that are optimal for the formation of

362

a strong 3D network. In addition, the smaller pores observed for the 20 g/L aerogels compared to

363

the 17.5 g/L aerogels would result in fewer CNFs in each cell wall which in turn also affects the

364

total strength of each wall.

365

Mechanical resilience

366

The ability of the aerogels to be resoaked, reshaped and redried suggests that the cell walls are

367

resistant to the mechanical deformation during the reshaping and can maintain their structure.

368

The mechanical properties of 17.5 g/L aerogels were therefore evaluated after several cycles of

369

dry compression, wet shape recovery, and drying. Figure 6(a) shows photographs of the shape

370

recovery of a compressed aerogel upon contact with water. The relative mechanical properties

371

and volume of the cycled aerogels, compared to the first cycle, are shown in Figure 6b.

ACS Paragon Plus Environment

20

Page 21 of 41 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

Biomacromolecules

372 373

Figure 6. (a) A sequence of photographs showing the shape recovery of a compressed aerogel,

374

(b) plot of the relative bulk modulus, yield stress and volume retention of a 17.5 g/L aerogel as a

375

function of consecutive compression, reswelling and drying cycles, (c) SEM image of the cross

376

section of an aerogel after 5 cycles of dry compression, reswelling and drying (scale bar 500 µm)

377

and (d) SEM image showing buckling and cracking (arrows) of the pore walls after 5

378

compression cycles (scale bar 1 µm)

379

The sequential compression, shape recovery and drying of the aerogels had no significant

380

effect on the density which remained almost constant, since the aerogels displayed a volumetric

ACS Paragon Plus Environment

21

Biomacromolecules 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 22 of 41

381

retention of above 90% even after 5 compression cycles. The modulus and yield strength on the

382

other hand were largely affected by the cycling, and dropped after one cycle, as can be seen in

383

Figure 6b. The modulus dropped to a value approximately 60% of the initial value and

384

subsequently remained stable during cycles 2-4 before dropping to around 30% of the initial

385

value in the fifth cycle. The yield stress displayed an initial drop to just over 20% of the initial

386

yield stress after which it increased in cycle 3 and remained stable in cycle 4, to approximately

387

55-60% of the initial, before it decreased in cycle 5. The overall decrease in mechanical

388

performance of the aerogels can be ascribed to the changes occurring in the structure during the

389

compression. Figure 6c shows that the overall pore structure changed compared to the initial

390

structure. The pores were significantly less regular and the pore walls were more disordered. The

391

examination of individual pore walls revealed a buckled surface, with some local cracks (Fig 6d).

392

In addition, the majority of the pore walls had creases on their surfaces, see Fig S3 in the

393

Supporting Information. The regain in yield stress and the stable value of the modulus can be

394

suggested to be due to the reformation of new adhesive contacts between the fibrils after each

395

cycle which is reasonable due to the close proximity of the fibrils in the cell wall. Since the

396

material becomes weaker after the mechanical deformation cycles and since water has been

397

shown to greatly plasticize hemiacetal crosslinked structures 22 it is possible that the material can

398

reorganize in the swollen state and subsequently reform new bonds when redried and hence show

399

self-healing properties. This is consistent with the reshaping and re-crosslinking of the

400

cylindrical aerogel into a prism shape where the cell walls can be reformed in the plasticized

401

state to adopt the new prism format when dried. This reshaping ability is however not unlimited

402

since the mechanical performance drops significantly after 5 cycles. This is supposedly due to

403

the cracks observed in the pore walls after 5 cycles and this is probably due to that the fibrils

ACS Paragon Plus Environment

22

Page 23 of 41 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

Biomacromolecules

404

have been permanently separated and no longer can form adhesive contacts as the fibrils have to

405

be in very close proximity to react and form hemiacetals.23

406 407

Functionalization of the aerogels using a biomimicking approach and possible applications Several structures found naturally such as cancellous bone 38, the interior of porcupine quills 39-

408 409

40

410

presented in the present work. All these natural materials also share a core shell structure where

411

the interior porous structure provides weight reduction and structural integrity while the outer

412

stiff layer provides the mechanical strength of the composite. A covering of the outer surface of

413

the aerogels with a stiff wax would therefore create a core-shell composite similar to the

414

structure that has been optimized by nature. The spin coating of molten carnauba wax on the

415

aerogels resulted in complete coverage of four adjoining sides of the porous aerogel. The average

416

thickness of the coated layer was 300 µm but variations in the structure of the aerogel surface

417

naturally resulted in variations of the wax coating layer. Since cellulose has well documented

418

amphiphilic nature, i.e. a high water wettability and a high dispersive surface energy 42-43 the wax

419

wets the pore walls and can therefore penetrate into the aerogel to different extents depending on

420

the local structure. The majority of the wax is however located on the surface where it sealed the

421

pores and created a solid wax crust (see Figure 7a). The penetration of the wax into the structure

422

is also indeed similar to how the keratin shell found in porcupine quills extends into the interior

423

of the quill39. The density of the wax coated aerogels increased fourfold from 31.8 kg/m3

424

(17.5 g/L aerogels) to 122.8 kg/m3, a value similar to that of a toucan beak (100 kg/m3) 41. The

425

effect of wax coating on the mechanical properties was tested in uniaxial compression and the

426

elastic modulus increased from approximately 1 MPa for the 17.5 g/L aerogel to 10 MPa for the

and the beak of the toucan bird

41

resemble the porous structure found in the aerogels

ACS Paragon Plus Environment

23

Biomacromolecules 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 24 of 41

427

corresponding coated aerogels. Figure 7b displays the initial part of the compression curves for a

428

wax coated and an uncoated aerogel. The tenfold increase in modulus is substantial considering

429

that the wax coating is only approximately 300 µm thick, which is comparable to the 500 µm

430

thick keratin layer of the toucan beak 41. The wax coating effectively increased the modulus and

431

the strength of the material and the composite displayed the dual effect of a porous interior and

432

stiff exterior. To exemplify the versatility of this core/shell aerogel, a toucan beak was moulded,

433

mechanically functionalized by a wax coating, see inset in Figure 7b, and spray painted to mime

434

the stiff and light weight shape and colour of a real toucan beak.

435 436

Figure 7 (a) SEM image of a cross section of a wax-coated aerogel (scale bar 250 µm), and (b)

437

Plot of the mechanical compressive stress vs. strain of reference and wax-coated aerogels, inset

438

is a photograph of a 16 cm long wax coated toucan beak inspired aerogel.

439

The porous inner structure of the aerogel also provides a large specific surface area that could

440

be utilized for further functionalization. By soaking the wet stable aerogels in a solution

441

containing pyrrole, a dopant and an initiator, a thin film of PPy was synthesized to cover the

442

entire aerogel structure. A linear relationship between the achieved PPy loading and the initial

443

pyrrole concentration was observed and aerogels with PPy loadings ranging from 11.3 wt%

ACS Paragon Plus Environment

24

Page 25 of 41 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

Biomacromolecules

444

(0.02 M Py) to 44 wt% (0.06 M Py) were prepared. The coating did not affect the external

445

dimensions of the aerogels which remained intact also after solvent exchange and drying under

446

ambient conditions. The polymerization of PPy turned the aerogels completely black and the

447

dark colour extended throughout the entire interior of the aerogel suggesting an even PPy

448

coverage. The coating itself consisted of closely packed PPy nanoparticles, displayed in

449

Figure 8a, with an average diameter of 35 nm which created an interconnected network verified

450

by the electrical conductivity of the aerogels. The electrical conductivity increased with PPy

451

loading up to a maximum value of 7 mS/cm for the aerogel containing 44 wt% PPy. The

452

conductivity value was calculated by considering the total aerogel volume including air. The

453

conductivity of the low density PPy composite was enough to light up a red LED connected by

454

two pieces of PPy coated aerogels exposed to a potential of 6V, see Figure 8b.

455

To further exemplify the versatility of crosslinked aerogel, a complex electrically conducting

456

solenoid, similar to that which was prepared from a CNF nanosheet with conducting polymers, 44

457

was prepared by reshaping a 50 cm long, 4 mm diameter aerogel cylinder, prior to

458

polymerisation of pyrrole on the surfaces within the porous structure. The entire solenoid had a

459

resistance of 7 kOhm and created an average magnetic field of 0.4 µT in the direction of the

460

solenoid by passing a current (3 mA) through the solenoid, see Figure 8c-d.

ACS Paragon Plus Environment

25

Biomacromolecules 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 26 of 41

461 462

Figure 8 (a) SEM image of the PPy coating on the surface of the functionalized aerogel (scale

463

bar 250 nm), photographs of (b) a lit LED connected by PPy-coated aerogels, (c) a PPy-

464

functionalized solenoid aerogel and the magnetometer (scale bar 5 mm), and (d) plot of the

465

magnetic field generated inside the solenoid as the current is turned on and off, the solid lines

466

represents the average magnetic field.

467

Conclusions

468

It has been shown that it is possible to mould, extrude, and reshape wet-stable CNF based

469

aerogels into complex shapes based on the freezing-induced crosslinking technique. Depending

470

on the shaping method CNF aerogels with densities ranging from 12 kg/m3 to 34 kg/m3 (7.5 g/L

471

– 20 g/L) could be obtained. All samples, independent of concentration, displayed a volumetric

472

retention of 75% after solvent exchange and ambient drying. The dispersions in the higher end of

ACS Paragon Plus Environment

26

Page 27 of 41 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

Biomacromolecules

473

the concentration range had a viscosity high enough to allow the periodate/CNF mixture to be

474

extruded through a nozzle and retain the shape in the liquid state before being crosslinked by

475

freezing. This opens up the possibility of using the freeze-linking method in combination with

476

3D-printing to produce unconventional wet-stable 3D structures. Additionally by re-soaking a

477

dry aerogel in water it became plasticized and could be reshaped into new macro structures such

478

as solenoids that are not possible to achieve by moulding. The formed macro structure became

479

locked and crosslinked into its new shape following either a solvent exchange or a freezing-

480

solvent exchange step which introduced new crosslinks between the CNFs.

481

The compressive modulus and yield strength scaled with an increasing density of the aerogels

482

formed at a concentration of up to 15 g/L after which the material property first increased

483

significantly for the samples formed at 17.5 g/L and subsequently decreased for the aerogels

484

formed at 20 g/L. The sharp increase and subsequent drop was attributed to the conditions in the

485

dispersion and interactions between the CNFs during freezing rather than simply the density or

486

the microscopic internal structure of the aerogels. At higher concentrations the CNF movements

487

became restricted earlier during freezing which inhibits an efficient crosslinking. The aerogels

488

showed shape memory effects exemplified by the 17.5 g/L dry aerogels that could be

489

compressed to 80%, re-soaked in water and re-dried 5 times without any significant shape or

490

volumetric loss. The mechanical performance was affected by the cycling and a decrease in

491

modulus and yield stress was observed after 1 cycle. However, the retained and in the case of

492

regained mechanical property during cycles 2-4 is argued to be due to the softening of the cell

493

walls when reswollen in water allowing the CNFs to rearrange and hence to allow for a renewed

494

crosslinking upon re-drying.

ACS Paragon Plus Environment

27

Biomacromolecules 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 28 of 41

495

One of the main advantages of aerogels is that both the inner porous structure and the outer

496

surface can be functionalized for different applications. The current work illustrates that it is

497

possible to completely coat the inner structure of the aerogel with the conducting polymer PPy

498

introducing enough conductivity to be able to light up an LED and to generate a 0.4 µT magnetic

499

field inside a functionalized solenoid shaped aerogel. Another example shows that by applying a

500

thin coating of carnauba wax to the surface of a toucan beak shaped aerogel, it is possible to

501

mimic a natural stiff and light weight structure and increase the modulus by a factor of ten. These

502

three examples show that it is possible to combine advanced shaping with functionalization of

503

both the inner structure and the surface of the aerogels, radically extending the possible use of

504

aerogels.

505

Associated content

506

SEM image of a freeze-dried 0.1 g/L oxidized CNF dispersion; photographs of a 7.5 g/L and a

507

17.5 g/L aerogel soaked in water; SEM image of the cross section of a 5 times cyclically

508

compressed aerogel displaying the creases observed on the surfaces of the 2D sheets comprised

509

of CNFs.

510

Author information

511

Corresponding Authors

512

*(J.E.) E-mail: [email protected]

513

*(L.W.) E-mail: [email protected]

514

Notes

515

The authors declare no competing financial interest.

516 517

ACS Paragon Plus Environment

28

Page 29 of 41 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

Biomacromolecules

518 519 520 521

Acknowledgements

522

Johan Erlandsson acknowledges the Swedish Energy Agency through the Modulit project

523

(grant number 37716-1) and Vinnova through the Digitial Cellulose Centre for financial support.

524

Hugo Françon and Hjalmar Granberg acknowledge Stiftelsen för Strategisk Forskning through

525

the 0D-3D project, Andrew Marais acknowledges Knut and Alice Wallenberg foundation for

526

financial support. Lars Wågberg acknowledges the Wallenberg Wood Science Centre for

527

financial support.

528

References

529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548

1. Siró, I.; Plackett, D.; Hedenqvist, M.; Ankerfors, M.; Lindström, T., Highly transparent films from carboxymethylated microfibrillated cellulose: The effect of multiple homogenization steps on key properties. J. Appl. Polym. Sci. 2011, 119, 2652-2660. 2. Belbekhouche, S.; Bras, J.; Siqueira, G.; Chappey, C.; Lebrun, L.; Khelifi, B.; Marais, S.; Dufresne, A., Water sorption behavior and gas barrier properties of cellulose whiskers and microfibrils films. Carbohydr. Polym. 2011, 83, 1740-1748. 3. Aulin, C.; Gällstedt, M.; Lindström, T., Oxygen and oil barrier properties of microfibrillated cellulose films and coatings. Cellulose 2010, 17, 559-574. 4. Mittal, N.; Jansson, R.; Widhe, M.; Benselfelt, T.; Håkansson, K. M. O.; Lundell, F.; Hedhammar, M.; Söderberg, L. D., Ultrastrong and Bioactive Nanostructured Bio-Based Composites. ACS Nano 2017, 11, 5148-5159. 5. Hamedi, M. M.; Hajian, A.; Fall, A. B.; Håkansson, K.; Salajkova, M.; Lundell, F.; Wågberg, L.; Berglund, L. A., Highly Conducting, Strong Nanocomposites Based on Nanocellulose-Assisted Aqueous Dispersions of Single-Wall Carbon Nanotubes. ACS Nano 2014, 8, 2467-2476. 6. Cervin, N.; Aulin, C.; Larsson, P.; Wågberg, L., Ultra porous nanocellulose aerogels as separation medium for mixtures of oil/water liquids. Cellulose 2012, 19, 401-410. 7. Cervin, N. T.; Johansson, E.; Larsson, P. A.; Wågberg, L., Strong, Water-Durable, and Wet-Resilient Cellulose Nanofibril-Stabilized Foams from Oven Drying. ACS Appl. Mater. Interfaces 2016, 8, 11682-11689.

ACS Paragon Plus Environment

29

Biomacromolecules 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

549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594

Page 30 of 41

8. Paakko, M.; Vapaavuori, J.; Silvennoinen, R.; Kosonen, H.; Ankerfors, M.; Lindstrom, T.; Berglund, L. A.; Ikkala, O., Long and entangled native cellulose I nanofibers allow flexible aerogels and hierarchically porous templates for functionalities. Soft Matter 2008, 4, 2492-2499. 9. Sehaqui, H.; Zhou, Q.; Berglund, L. A., High-porosity aerogels of high specific surface area prepared from nanofibrillated cellulose (NFC). Compos. Sci. Technol. 2011, 71, 1593-1599. 10. Yang, X.; Cranston, E. D., Chemically Cross-Linked Cellulose Nanocrystal Aerogels with Shape Recovery and Superabsorbent Properties. Chem. Mater. 2014, 26, 6016-6025. 11. Sehaqui, H.; Salajkova, M.; Zhou, Q.; Berglund, L. A., Mechanical performance tailoring of tough ultra-high porosity foams prepared from cellulose I nanofiber suspensions. Soft Matter 2010, 6, 1824-1832. 12. Martoïa, F.; Cochereau, T.; Dumont, P. J. J.; Orgéas, L.; Terrien, M.; Belgacem, M. N., Cellulose nanofibril foams: Links between ice-templating conditions, microstructures and mechanical properties. Materials & Design 2016, 104, 376-391. 13. Saito, T.; Kimura, S.; Nishiyama, Y.; Isogai, A., Cellulose Nanofibers Prepared by TEMPO-Mediated Oxidation of Native Cellulose. Biomacromolecules 2007, 8, 2485-2491. 14. Wågberg, L.; Decher, G.; Norgren, M.; Lindström, T.; Ankerfors, M.; Axnäs, K., The Build-Up of Polyelectrolyte Multilayers of Microfibrillated Cellulose and Cationic Polyelectrolytes. Langmuir 2008, 24, 784-795. 15. Hamedi, M.; Karabulut, E.; Marais, A.; Herland, A.; Nyström, G.; Wågberg, L., Nanocellulose Aerogels Functionalized by Rapid Layer-by-Layer Assembly for High Charge Storage and Beyond. Angew. Chem. Int. Ed. 2013, 52, 12038-12042. 16. Kim, C.; Youn, H.; Lee, H., Preparation of cross-linked cellulose nanofibril aerogel with water absorbency and shape recovery. Cellulose 2015, 10.1007/s10570-015-0745-5, 1-10. 17. Zhang, W.; Zhang, Y.; Lu, C.; Deng, Y., Aerogels from crosslinked cellulose nano/micro-fibrils and their fast shape recovery property in water. J. Mater. Chem. 2012, 22, 11642-11650. 18. Cai, H.; Sharma, S.; Liu, W.; Mu, W.; Liu, W.; Zhang, X.; Deng, Y., Aerogel Microspheres from Natural Cellulose Nanofibrils and Their Application as Cell Culture Scaffold. Biomacromolecules 2014, 15, 2540-2547. 19. Köklükaya, O.; Carosio, F.; Wågberg, L., Superior Flame-Resistant Cellulose Nanofibril Aerogels Modified with Hybrid Layer-by-Layer Coatings. ACS Appl. Mater. Interfaces 2017, 10.1021/acsami.7b08018. 20. Henschen, J.; Illergård, J.; Larsson, P. A.; Ek, M.; Wågberg, L., Contact-active antibacterial aerogels from cellulose nanofibrils. Colloids and Surfaces B: Biointerfaces 2016, 146, 415-422. 21. Nystrom, G.; Marais, A.; Karabulut, E.; Wagberg, L.; Cui, Y.; Hamedi, M. M., Selfassembled three-dimensional and compressible interdigitated thin-film supercapacitors and batteries. Nat Commun 2015, 6. 22. Erlandsson, J.; López Durán, V.; Granberg, H.; Sandberg, M.; Larsson, P. A.; Wågberg, L., Macro- and mesoporous nanocellulose beads for use in energy storage devices. Applied Materials Today 2016, 5, 246-254. 23. Erlandsson, J.; Pettersson, T.; Ingverud, T.; Granberg, H.; Larsson, P. A.; Malkoch, M.; Wågberg, L., On the mechanism behind freezing-induced chemical crosslinking in ice-templated cellulose nanofibril aerogels. J. Mater. Chem. A 2018, 6, 19371-19380. 24. Katz, S.; Beatson, R.; Scallan, A. M., The determination of strong and weak acidic groups in sulfite pulps. Sven. Papperstidn. 1984, R48-R53.

ACS Paragon Plus Environment

30

Page 31 of 41 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

595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638

Biomacromolecules

25. Zhao, H.; Heindel, N., Determination of Degree of Substitution of Formyl Groups in Polyaldehyde Dextran by the Hydroxylamine Hydrochloride Method. Pharm Res 1991, 8, 400402. 26. Munier, P.; Gordeyeva, K.; Bergström, L.; Fall, A. B., Directional Freezing of Nanocellulose Dispersions Aligns the Rod-Like Particles and Produces Low-Density and Robust Particle Networks. Biomacromolecules 2016, 17, 1875-1881. 27. Wicklein, B.; Kocjan, A.; Salazar-Alvarez, G.; Carosio, F.; Camino, G.; Antonietti, M.; Bergström, L., Thermally insulating and fire-retardant lightweight anisotropic foams based on nanocellulose and graphene oxide. Nat Nano 2015, 10, 277-283. 28. Deville, S.; Adrien, J.; Maire, E.; Scheel, M.; Di Michiel, M., Time-lapse, threedimensional in situ imaging of ice crystal growth in a colloidal silica suspension. Acta Mater. 2013, 61, 2077-2086. 29. Erlandsson, J.; Pettersson, T.; Ingverud, T.; Granberg, H.; Larsson, P. A.; Malkoch, M.; Wagberg, L., On the mechanism behind freezing-induced chemical crosslinking in ice-templated cellulose nanofibril aerogels. J. Mater. Chem. A 2018, 10.1039/C8TA06319B. 30. Li, W. L.; Lu, K.; Walz, J. Y., Freeze casting of porous materials: review of critical factors in microstructure evolution. Int. Mater. Rev. 2012, 57, 37-60. 31. Wu, S.; Zhu, C.; He, Z.; Xue, H.; Fan, Q.; Song, Y.; Francisco, J. S.; Zeng, X. C.; Wang, J., Ion-specific ice recrystallization provides a facile approach for the fabrication of porous materials. Nat. Commun. 2017, 8, 15154. 32. Naderi, A.; Lindström, T.; Pettersson, T., The state of carboxymethylated nanofibrils after homogenization-aided dilution from concentrated suspensions: a rheological perspective. Cellulose 2014, 21, 2357-2368. 33. Chen, W.; Li, Q.; Wang, Y.; Yi, X.; Zeng, J.; Yu, H.; Liu, Y.; Li, J., Comparative study of aerogels obtained from differently prepared nanocellulose fibers. ChemSusChem 2014, 7, 15461. 34. Osorio, D. A.; Seifried, B.; Moquin, P.; Grandfield, K.; Cranston, E. D., Morphology of cross-linked cellulose nanocrystal aerogels: cryo-templating versus pressurized gas expansion processing. J. Mater. Sci. 2018, 53, 9842-9860. 35. Guigo, N.; Mazeau, K.; Putaux, J.-L.; Heux, L., Surface modification of cellulose microfibrils by periodate oxidation and subsequent reductive amination with benzylamine: a topochemical study. Cellulose 2014, 21, 4119-4133. 36. Kim, U.-J.; Kuga, S.; Wada, M.; Okano, T.; Kondo, T., Periodate Oxidation of Crystalline Cellulose. Biomacromolecules 2000, 1, 488-492. 37. Gibson, L. J., Cellular solids : structure & properties / Lorna J. Gibson, Michael F. Ashby. Pergamon Press: Oxford [Oxfordshire] ; New York, 1988. 38. Meyers, M. A.; Chen, P.-Y.; Lin, A. Y.-M.; Seki, Y., Biological materials: Structure and mechanical properties. Prog. Mater Sci. 2008, 53, 1-206. 39. Yang, W.; Chao, C.; McKittrick, J., Axial compression of a hollow cylinder filled with foam: A study of porcupine quills. Acta Biomater. 2013, 9, 5297-5304. 40. Yang, W.; McKittrick, J., Separating the influence of the cortex and foam on the mechanical properties of porcupine quills. Acta Biomater. 2013, 9, 9065-9074. 41. Seki, Y.; Kad, B.; Benson, D.; Meyers, M. A., The toucan beak: Structure and mechanical response. Mater. Sci. Eng., C 2006, 26, 1412-1420.

ACS Paragon Plus Environment

31

Biomacromolecules 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

639 640 641 642 643 644 645 646

Page 32 of 41

42. Lehtiö, J.; Sugiyama, J.; Gustavsson, M.; Fransson, L.; Linder, M.; Teeri, T. T., The binding specificity and affinity determinants of family 1 and family 3 cellulose binding modules. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 484-489. 43. Lindman, B.; Medronho, B.; Alves, L.; Costa, C.; Edlund, H.; Norgren, M., The relevance of structural features of cellulose and its interactions to dissolution, regeneration, gelation and plasticization phenomena. PCCP 2017, 19, 23704-23718. 44. Malti, A.; Tu, D.; Edberg, J.; Sani, N.; Rudd, S.; Evans, D.; Forchheimer, R., Electromagnetic devices from conducting polymers. Org. Electron. 2017, 50, 304-310.

647

648

ACS Paragon Plus Environment

32

Page 33 of 41 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

Biomacromolecules

88x36mm (300 x 300 DPI)

ACS Paragon Plus Environment

Biomacromolecules 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

Figure 1 (a) Schematic representation of the preparation of the aerogels, (b) shaping by moulding or extrusion (c) post production reshaping of aerogels, (d) schematic view of the ice crystals excluding the CNFs from the ice-phase and tightly packing them together in the lamellae.

ACS Paragon Plus Environment

Page 34 of 41

Page 35 of 41 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

Biomacromolecules

Figure 2 (a) Photograph of a spherical cross-linked CNF aerogel made from 20 g/L CNF gel (scale bar = 1 cm), (b) SEM-image of the cross-section of the spherical aerogel (scale bar = 500 µm), (c) SEM image showing the pore walls (scale bar = 10 µm), (d) The aerogel density as a function of the initial CNF concentration. 96x139mm (300 x 300 DPI)

ACS Paragon Plus Environment

Biomacromolecules 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

Figure 3 Photographs of (a) an extruded swirl aerogel, (b) a solenoid- shaped aerogel prepared by reshaping a long cylindrical aerogel , (d) a cylindrical aerogel (left) reshaped into a prism (right). The scale bars are 1 cm. 150x128mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 36 of 41

Page 37 of 41 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

Biomacromolecules

Figure 4. SEM images of cross sections of moulded (cubic) aerogels prepared from (a) 7.5 g/L, (b) 17.5 g/L and (c) 20 g/L. Scale bars represent 1 mm. 384x88mm (300 x 300 DPI)

ACS Paragon Plus Environment

Biomacromolecules 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

Figure 5. (a) Compressive stress strain curves for the freeze-linked aerogels prepared from different initial CNF concentrations. The inset shows the initial part of the compression test and (b) the cell wall properties calculated using the Ashby and Gibson scaling laws. 292x108mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 38 of 41

Page 39 of 41 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

Biomacromolecules

Figure 6. (a) A sequence of photographs showing the shape recovery of a compressed aerogel, (b) plot of the relative bulk modulus, yield stress and volume retention of a 17.5 g/L aerogel as a function of consecutive compression, reswelling and drying cycles, (c) SEM image of the cross section of an aerogel after 5 cycles of dry compression, reswelling and drying (scale bar 500 µm) and (d) SEM image showing buckling and cracking (arrows) of the pore walls after 5 compression cycles (scale bar 1 µm) 164x190mm (300 x 300 DPI)

ACS Paragon Plus Environment

Biomacromolecules 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

Figure 7 (a) SEM image of a cross section of a wax-coated aerogel (scale bar 250 µm), and (b) Plot of the mechanical compressive stress vs. strain of reference and wax-coated aerogels, inset is a photograph of a 16 cm long wax coated toucan beak inspired aerogel.

ACS Paragon Plus Environment

Page 40 of 41

Page 41 of 41 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

Biomacromolecules

Figure 8 (a) SEM image of the PPy coating on the surface of the functionalized aerogel (scale bar 250 nm), photographs of (b) a lit LED connected by PPy-coated aerogels, (c) a PPy-functionalized solenoid aerogel and the magnetometer (scale bar 5 mm), and (d) plot of the magnetic field generated inside the solenoid as the current is turned on and off, the solid lines represents the average magnetic field. 180x136mm (300 x 300 DPI)

ACS Paragon Plus Environment