Tunable Structural and Mechanical Properties of ... - ACS Publications

May 9, 2017 - Macromolecules Innovation Institute, Virginia Tech, Department of Materials Science & Engineering, 445 Old Turner Street, 203 Holden Hal...
0 downloads 9 Views 2MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Tunable Structural and Mechanical Properties of Cellulose Nanofiber Substrates in Aqueous Conditions for Stem Cell Culture Megan Smyth, Carole Fournier, Carlos Driemeier, Catherine Picart, E. Johan Foster, and Julien Bras Biomacromolecules, Just Accepted Manuscript • Publication Date (Web): 09 May 2017 Downloaded from http://pubs.acs.org on May 10, 2017

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 free 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 accessible to all readers and 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.

Biomacromolecules 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 37

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

Tunable Structural and Mechanical Properties of Cellulose Nanofiber Substrates in Aqueous

2

Conditions for Stem Cell Culture

3

Megan Smyth1,2, Carole Fournier3,4, Carlos Driemeier5, Catherine Picart3,4 E. Johan Foster6*,

4

and Julien Bras1,2,7*

5

1. CNRS, LGP2, 461 Rue de la Papeterie, 38402, Saint-Martin-d'Hères, France

6

2. Université Grenoble Alpes, LGP2, 38000 Grenoble, France

7

3. CNRS, UMR 5628, LMGP, 38016 Grenoble, France

8

4. Université Grenoble Alpes, Grenoble Institute of Technology, 38016 Grenoble, France

9

5. Centro Nacional de Pesquisa em Energia e Materiais (CNPEM), Laboratório Nacional de

10

Ciência e Tecnologia do Bioetanol (CTBE), 13083-970 Campinas, São Paulo, Brazil

11

6. Virginia Polytechnic Institute and State University (Virginia Tech), Macromolecules

12

Innovation Institute (MII), Department of Materials Science and Engineering, Blacksburg,

13

Virginia 24061, USA

14

7. Institut Universitaire de France, 75005 Paris, France

15 16

KEYWORDS: Cell adhesion, cell viability, nanofibrillated cellulose, microfibrillated cellulose,

17

mechanical properties, material properties

18 19 20

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

21

ABSTRACT

22

Thin cellulose nanofiber (CNF) nanostructured substrates with varying roughness, stiffness

23

(Young’s modulus), porosity, and swelling properties were produced, by varying the conditions

24

used during fabrication. It was shown that with increased heat exposure, CNF substrate porosity

25

in an aqueous state decreased while Young’s modulus in a water submerged state increased. In

26

this study, the adhesion and viability of mesenchymal stem cells (MSCs) cultured on this CNF

27

substrate will be presented. Viability of D1/BALBc MSCs were assessed for 24 and 48 hours,

28

and it was shown that depending on the CNF substrate the viability varied significantly. The

29

adhesion of MSCs after 6 and 24 hours was conditional on material mechanical properties and

30

porosity of the CNF in cell culture conditions. These results suggest that material properties of

31

CNF nanostructured substrate within the aqueous state can be easily tuned with curing step

32

without any chemical modification to the CNF, and that these changes can affect MSC viability

33

in cell culture.

34

Table of Contents Graphic

35 36

INTRODUCTION

37

With being the most abundant polymer on earth, interest in the use of cellulose for a variety of

38

applications from paper or polymer composites to biomedical applications has increased annually

ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37

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

39

since the 1980s 1. Cellulose has favorable properties for biomedical researchers such as

40

biocompatibility, renewability and low cytotoxicity 2. The use of cellulose in biomedical

41

applications such as drug delivery, tissue engineering, and wound healing has become more of an

42

interest to researchers who want to use natural polymers 3. Since the 2000’s, a special focus on

43

nanoscale cellulose has been observed for such applications. For example, bacterial nanocellulose

44

(BNC) is commonly used in biomedical applications and is readily available on the market,

45

especially in the field of wound healing 4. Unfortunately, such material is expensive, its

46

production is difficult to upscale since such membranes are processed with difficulties. The use of

47

other forms of nanoscaled cellulose, cellulose nanocrystals (CNC) or microfibrillated cellulose

48

(MFC also called CNF), has gained greater interest with the announcement of their

49

industrialization the last 5 years. Respectively, discovered in the 50’s 5 and the 80’s 6 , they are

50

obtained from any cellulose fibers7 either by a chemical hydrolysis for CNC or by a mechanical

51

disintegration, mostly after a pre-treatment, for CNF. Books 8 and reviews 9-10 details their

52

production 11 and properties 12-14, but very recently they have been used in the field of biomedical

53

engineering with a focus on using cellulose as a scaffold for tissue engineering or cell culture 2, 15.

54

CNCs are mostly used to impart specific chemical properties to promote cell adhesion and direct

55

cell growth 16,17 or to provide a structure for cell adhesion in a particular direction 18. CNF has

56

been used as a tissue engineering scaffold in as an additive to a matrix, and most commonly

57

hydrogels 19,20,21,22. Other processing techniques can be used to create celluose based substrates

58

such as electrospinning23-24 and areogels25. Both of these techniques have been employed in

59

biomedical applications. It is worth noting that most of these scientific papers are very recent

60

(less than 2-3 years), proving by itself the novelty of our study.

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

Page 4 of 37

61

Indeed microfibrillated cellulose has been first extensively studied for its use in composites and

62

as barrier films 26,12. Compared to cellulose nanocrystals, CNF is less crystalline, flexible and has

63

diameter of approximately 20-60 nm and length of several microns 12. The morphological

64

properties of CNF can vary depending on the source of the CNF, and the means of production.

65

Enzymatic pre-treatment, mechanical fibrillation, homogenization, and chemical treatment are co

66

mmon ways to produce CNF. Each production technique imparts different characteristics to the

67

CNF, such as surface groups, charge, and crystallinity 11. CNF has numerous hydrogen bonds on

68

the surface of the fiber, and because of these bonds and the entanglement of these flexible

69

nanofibers, they can produce a nanostructured substrate, also called films or nanopaper. Such

70

substrates present very good barrier or mechanical properties and can also be transparent. There

71

are different means to produce such nanopaper 27-28, and their large scale production process is

72

still an issue in spite of announcement of pilot scale roll of such films from VTT. Depending on

73

the manufacturing procedure, their properties might change as recently shown for barriers

74

applications 29. Others have shown that the structural properties, which influence barrier

75

properties of the CNF film might differ based on drying techniques 29-31, but there has been no

76

research conducted to fine-tune the mechanical properties for biomedical applications.

77

The use of mesenchymal stem cells was first discovered by Friedenstein in the 1970s

78

research on the work of Friedenstein showed that cells isolated from bone marrow had the ability

79

to differentiate into osteoblasts, adipocytes, and chondrocytes

80

many factors including adhesion and proliferation. It has been shown that without adhesion stem

81

cells will undergo apoptosis, since stem cells are reliant on substrate attachment, or are anchorage

82

dependent

83

microenvironment leading to contractility and reorganization of the actin filaments within the

34-35

33

32

. Further

. MSC viability is dependent on

. This anchorage of the cells at the cellular level occurs when cells probe their

ACS Paragon Plus Environment

Page 5 of 37

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

35-36

84

cytoskeleton into filopodia

. The adhesion of stem cells can be dependent on various factors

85

such as surface chemistry 37, stiffness

86

Engler and colleagues it was shown that MSCs can present different morphologies and spreading

87

patterns dependent on substrate stiffness, and that these differences in morphology were indicative

88

of various differentiation pathways 39. The variation of the mechanical properties of polymer based

89

substrates such as polyacrylamide 39 and the effect on MSC adhesion and viability has been studied

90

extensively, but to the best of the authors’ knowledge never on nanocellulose based materials with

91

tunable material properties.

92

Culturing of MSCs onto cellulose based substrates has been performed by a variety of

93

researchers. Most of the work is recent and focuses on bacterial nanocellulose substrates, not

94

CNF produced from wood, which is the focus of this study. Krontiras and colleagues used murine

95

MSCs for adipogenic differentiation using bacterial nanocellulose and alginate scaffolds. It was

96

shown that after 4 weeks in culture, adipocytes displayed lipid formation 40. BNC has been used

97

alone as a scaffold in work by Favi et al. Favi showed that BNC scaffolds seeded with equine

98

MSCs and differentiated in vitro promoted cell viability, proliferation, adhesion, and supported

99

chondrogenic and osteogenic differentiation 41. For plant-based nanocellulose, limited work has

35

, and topography

34, 36, 38

. In seminal work conducted by

100

been completed using MSCs with both CNCs and CNF. CNCs were dispersed in a poly(lactic

101

acid), followed by electrospinning to create a scaffold, and tested with human MSCs for basic

102

cytocompatibility by Zhou et al. 42. Xing and colleagues created a CNF hydrogel with gelatin and

103

tested with human MSCs and determined that MSCs maintained differentiation potential and

104

secreted extracellular matrix, but did not form cellular colonies on the CNF 19. Even if others

105

have checked the cytotoxicity of purely CNF aerogels or other substrates 22, 43 or performed

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

106

human cell culture 25, to the best of the authors’ knowledge, there has been no reported research

107

on only CNF based substrates with tunable mechanical properties and mesenchymal stem cells.

108

Thin CNF substrate mechanical and material properties were easily tuned by exposing the solvent

109

casted CNF films to elevated temperatures. This facile curing step created a change in the

110

material properties without any complicated chemical processes, or the addition of a polymer

111

matrix or filler. Mechanical and material measurements were analyzed in cell culture like

112

conditions in order to have a better understanding how the properties of CNF thin films are

113

effected by hydration and how these properties could impact future cell based studies. The

114

influence of surface roughness, porosity, and Young’s modulus on MSC adhesion and viability

115

will be reported.

116 117

MATERIALS AND METHODS

118

Materials

119

A 2 % wt suspension of enzymatically pre-treated CNF from wood pulp (Domsjo) was purchased

120

from Centre Technique du Papier (CTP) (France). Phosphate buffer solution (PBS), Minimum

121

Essential Medium Eagle (αMEM), 4,6-diaminidino-2-phenylindole-dilactate (DAPI),

122

phalloidintetramethylrhod B isothiocyanate dyes (Phalloidin), glycerol, mowiol 4-88, tris buffer

123

saline (TBS), thiazolyl blue tetrazolium bromide (MTT), and dimethyl sulfoxide (DMSO) were

124

purchased from Sigma-Aldrich (USA). BALB/c D1 murine bone marrow mesenchymal stem

125

cells were purchased from American Type Culture Collection (ATCC, France). 10% heat-

126

inactivated fetal bovine serum (FBS) was acquired from PAA Laboratories (Austria), and

ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37

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

127

penicillin streptomycin mixture and L-Glutamine were purchased from Invitrogen (USA). All

128

materials were used as delivered without any chemical modification.

129

Production of CNF Films

130

CNF was characterized to determine the cellulose content following the protocol of Rowell et

131

al.44, and zeta potential was measured a Malvern Zetasizer 2000 (Malvern, United Kingdom). 10

132

mL of approximately 0.001% suspension of CNF was measured in triplicate. Results are reported

133

in mean ± standard deviation. CNF membranes were produced by casting 1% wt solution of CNF

134

onto teflon mold and allowed to dry at ambient conditions for 5 days (AC-CNF) followed by

135

curing for 1 (1hC-CNF) or 2 (2hC-CNF) hours at 150°C following a modified protocol by Bardet

136

et al. 29, 31. Basis weight was determined by the following equation:

137

𝑔

𝑀𝑎𝑠𝑠 𝑜𝑓 𝐶𝑁𝐹 𝑓𝑖𝑙𝑚

𝐵𝑎𝑠𝑖𝑠 𝑊𝑒𝑖𝑔ℎ𝑡 (𝑚2 ) = 𝐴𝑟𝑒𝑎 𝑜𝑓 𝐶𝑁𝐹 𝑓𝑖𝑙𝑚 (Eq. 1)

138

Atomic Force Microscopy (AFM) Imaging

139

Images of the surface of CNF films were acquired using tapping mode with a Nanoscope IIA

140

microscope from Vecco Instruments (USA) at a frequency of ~265-340 Hz. Silicon cantilevers

141

with a curvature of 10-15 nm were used (OTESPA®, Bruker, USA). For surface roughness

142

measurements 10x10 µm images were acquired to analyze the surface roughness. All

143

measurements were analyzed using Nanoscope Analysis (USA). ImageJ software (NIH) was used

144

to analyze materials dimensions and at least 50 measurements have been performed on at least 5

145

different images and averages are presented.

146

Scanning Electron Microscopy (SEM) Imaging

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

147

SEM images of the CNF films for thickness measurements were obtained using a Quanta200®

148

(Netherlands). Samples were attached to a holder, coated with a thin layer of Au/Pb to create a

149

conductive charge, and scanned using an accelerated voltage of 10 kV and a working distance of

150

10.3 mm. An Everhart Thornley-Secondary electron detector (ETD) was employed. For thickness

151

measurements at least 25 measurements were taken from each sample using ImageJ software

152

(U.S National Institutes of Health, USA).

153

Thermoporometry

154

Calorimetric thermoporometry was performed following the protocol developed by Driemeier et

155

al. 45. Thermoporometry employed a differential scanning calorimeter DSC-Q200 (TA

156

Instruments, USA) with a RCS90 (TA Instruments, USA) cooling unit and an autosampler. CNF

157

samples were pre-conditioned by soaking in deionized water (DI H2O) for 24 hours. The water-

158

saturated samples were inserted into aluminum Tzero® pans later sealed with hermetic lids. The

159

DSC temperature program freezes samples to -70°C followed by 18 heating steps, each one made

160

of a heating ramp (1°C/min) followed by an equilibrating isotherm. Measured calorimetric signal

161

is analyzed to quantify the amount of ice melting below 0°C, termed Freezing Bound Water

162

(FBW). The temperature depression ΔT of ice melting is related to pore diameter d through the

163

Gibbs-Thomson equation, d = -2 Kc/ΔT, with Kc = 19.8 nm/K. The reported thermoporometry

164

profiles are the cumulative distributions of FBW (given in units of g water per g dry matter) as

165

function of d in the 1-200 nm range of pore size.

166

Mechanical Tests

167

Tensile tests and dynamic mechanical analysis (DMA) were measured using a TA Instruments

168

RSA 3 (USA) with a submersion clamp system designed internally. CNF films were tested in

ACS Paragon Plus Environment

Page 8 of 37

Page 9 of 37

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

169

physiological-like conditions, i.e. in aqueous environment. For physiological-like conditions,

170

CNF films were conditioned for 24 hours in PBS prior to mechanical tests. Samples underwent

171

rectangular tension measurements and compression testing. The distance between the clamps for

172

the tension was 10 mm, and the films had a width of approximately 5 mm and thickness of 0.2

173

mm. Samples underwent rectangular tension measurements at a rate of 0.001 mm/s at 37°C. For

174

compression, films were tested with an angular frequency from 0-20 Hz at 37°C with a constant

175

strain of 10%. Samples had dimensions of 15 mm in diameter and thickness of approximately 0.2

176

mm. The stress vs strain and frequency dependent curves were measured and analyzed using TA

177

Orchestrator (USA). Results were done in triplicate.

178

Cell Culture

179

Cell culture studies were performed with BALB/c D1 murine bone marrow mesenchymal stem

180

cells (D1 MSCs) (American Type Culture Collection, ATCC, France). Cells were cultured in

181

αMEM supplemented with 10% heat-inactivated fetal bovine serum, 1% penicillin streptomycin

182

mixture and 2 mM L-Glutamine. CNF films were sterilized with UV exposure prior to cell

183

seeding and rinsed twice with DI H2O. Cells were seeded at a density of 20,000 cells/cm2 for 6

184

hours and 24 hours for DAPI-phalloidin assay, and 24 hours and 48 hours for MTT analysis in a

185

humidified atmosphere of 5% CO2 at 37 °C.

186

DAPI-Phalloidin Assay

187

4,6-Diaminidino-2-phenylindole-dilactate (DAPI, 20 mg/mL, Sigma-Aldrich, USA) and

188

phalloidintetramethylrhod B isothiocyanate dyes (phalloidin, 10 mg/mL, Sigma-Aldrich,USA)

189

were used to determine the adhesion of cells at 6 hours and 24 hours following seeding on CNF

190

substrates and a glass control. At each time point, the culture media was moved and the samples

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

Page 10 of 37

191

were fixed with 3.7% paraformaldeyde. The paraformaldehyde was removed after 20 min at room

192

temperature and washed twice with PBS. The phalloidin/DAPI solution was added for 45 min at

193

room temperature (12.5 µg/mL phalloidin, 10 µg/mLDAPI). The samples were then washed three

194

times with PBS, mounted with anti-fade reagent, which is 24% glycerol, 9.6% mowiol 4-88 in

195

tris buffered saline (TBS). Followed by analysis by inverted fluorescent microscopy (Zeiss,

196

Germany) or confocal laser scanning microscopy (Zeiss, Germany) 24 hours later. The cell

197

surface area and aspect ratio were analyzed using ImageJ with a sample size of 150 cells per

198

condition using at least 12 different microscopy images from 4 samples from two independent

199

experiments. Cell surface area and aspect ratio measurements were done by measuring clearly

200

defined individual cells using the tracing tool in ImageJ after calibrating the scale and

201

thresholding each image.

202

MTT Assay

203

Thiazolyl blue tetrazolium bromide (MTT, 2.5 mg/mL, Sigma Aldrich, USA) was used to

204

perform a MTT assay on five independent experiments to measure the mitochondrial activity of

205

D1 MSCs after 24 hours and 48 hours post seeding on CNF substrates or tissue culture polysterne

206

(TCPS) control. MTT solution was added to the cells and incubated in media for 1 hour. After 1

207

hour, the growth media was removed and 500 µL of dimethyl sulfoxide (DMSO) was added to

208

solubilize the formazan crystals formed in the reduction of MTT. 100 µL of the formazan

209

solution was added to a microplate reader and diluted with 100 µL of DMSO. The absorbance of

210

each sample was read at 550 nm. The samples were imaged using optical microscopy (Ziess,

211

Germany).

212

Statistical Analysis

ACS Paragon Plus Environment

Page 11 of 37

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

213

All experiments were carried out in triplicate unless otherwise noted. All results are presented

214

with mean ± standard deviation unless otherwise noted. The R2 value of thermoporosity vs.

215

Young’s modulus was performed by Originlab using a linear-regression test. Statistical analysis

216

of cell studies (DAPI-Phalloidin Assay/MTT Assay) were performed in Originlab using a

217

Kolmogorov-Smirnov test for normality. For the cell area and aspect ratio results following the

218

DAPI-Phalloidin assay, the samples did not follow a normal distribution and were normalized

219

using a Box-Cox transformation. Following the normalization transformation, a Kolmogorov-

220

Smirnov test was repeated. Normalized results were analyzed by two-way ANOVA with post-hoc

221

testing using a Bonferroni test.

222 223

RESULTS AND DISCUSSION

224 225

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

226

Figure 1. Schematic of the experimental set-up. From raw material of CNF suspension to CNF

227

films and sequential curing of films and eventual cellular testing of CNF films for cell attachment

228

and viability

229 230

Proof of Concept - CNF to Cells

231

CNF used in this study had a diameter of 14 ± 5 nm as reported by Bardet et al.29, a cellulose

232

content of 86%, which is in accordance with literature 25, and a zeta potential of -17.0 ± 0.7 mV.

233

To determine the optimal properties of the CNF films before cell adhesion and viability tests,

234

several CNF films with varying basis weight was produced and either left as-castor dried at

235

150°C for 2 hours. Based on literature29, such curing treatment modifies the structures of this

236

entangled network of CNF. Others called it densification by hornification due to the increase in

237

number of strong hydrogen bonds and by the elimination of water. The mechanical properties of

238

these films were tested using traction testing and compression testing in aqueous conditions, i.e.

239

cell culture like conditions, 37°C submerged in PBS. The results from this work are presented in

240

Figure 2. As shown in the table, there are different mechanical properties of the films depending

241

on the basis weight and the Young’s modulus varies in regards to curing the films. The general

242

trend that with curing the films had a higher modulus compared to the as casted films.

ACS Paragon Plus Environment

Page 12 of 37

Page 13 of 37

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

243 244

Figure 2. A) Basis weight (g/m2), elongation at break (%), Young’s modulus (MPa), and storage

245

modulus (MPa) of CNF films as-cast(AC) or after 2h curing (2hC) following mechanical testing

246

at physiological conditions at 37°C in PBS. D1 MSCs interactions with the CNF with different

247

basis weight were assessed after 24 h of culture by staining the cell nucleus and cytoskeleton

248

(DAPI and phalloidin). B)1-AC-CNF, C) 2-2hC-CNF, D) 3-2hC-CNF, and E) 4-2hC-CNF as

249

measured by confocal laser scanning microscopy.

250 251

It is worth noting that with similar raw material, a significant difference in mechanical properties

252

(E’ in PBS is between 6 and 90 MPa) can result. This kind of ability to vary the properties is not

253

so common, even within the CNF field, with only few scientific papers dealing with such an

254

approach29, 31.

255

It is already recognized that mechanical values are crucial for cell growth and differences might

256

be interesting for stem cell differentiation39 which lead to the preliminary tests to determine if it

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

257

was possible to culture D1 MSCs with these different CNF substrates. D1 MSCs were cultured

258

for 24 hours, stained with Phalloidin/DAPI and observed using confocal microscopy as shown in

259

Figure 2. Thinner films allowed imaging and showed cell attachment for the high mechanical

260

property substrates, but low attachment for the as-cast films with the same fiber amount. This

261

result is promising because of the apparent attachment of the cells to the CNF, but it is difficult to

262

say if this qualitative difference is due to mechanical properties or to other substrate properties.

263 264

Tuning the Nanostructure Organization of CNF Substrate

265

Starting from this proof of concept, new CNF films were produced with same mass of fiber, but

266

the drying time was varied for tuning the intrinsic properties of such CNF films. The table below

267

summarizes the properties of such substrates.

ACS Paragon Plus Environment

Page 14 of 37

Page 15 of 37

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

Biomacromolecules

1

Sample

Transmittance Surface Freezing Bound Elongation Water Content PBS Uptake Basis Weight (%) 300-800 Roughness Water at < at Break 3 4 2 5 (%) (w/w) (%) (w/w%) (g/m ) nm1 (RMS) (nm)2 200 nm (g/g)6 (%)7

Young’s Modulus (E) (MPa)7

Storage Modulus (E’) (MPa) from 0.05-20 Hz8

0.022± 0.003 – AC-CNF

4.0-10.1

4.71±1.55

13.1±4.0

103.5±2.8

38.6±1.1

0.44± 0.006 4.2 ± 1.3

5.6 ± 0.6 0.305±0.120

1hC-CNF

3.8 -11.5

4.71±1.53

6.6±1.8

70.6±9.5

29.5±3.1

0.38± 0.02

52.2 ±

0.026 ± 0.011 –

26.3

0.329 ± 0.085

3.4 ± 1.3

0.050± 0.028 – 2hC-CNF

2.0 -8.2

4.34±0.48

3.4±1.4

65.3±3.7

28.1±2.8

0.36± 0.02

2.5 ± 0.6

116 ± 29.1 0.342 ± 0.194

2

Table 1. Summary of CNF films properties as measured by 1) UV spectroscopy (S1), 2) Root Mean Squared Roughness values from

3

AFM images of CNF films, as measured by nanoscope analysis AFM, 3) drying of films at 105°C for 24 hours (S2), 4) submersion of

4

films for 24 hours in PBS at 37°C (S2), 5) Basis Weight of CNF Films, 6) thermoporometry, 7) tensile testing at 37°C in PBS, 8)

5

compression DMA at 37°C in PBS.

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

6

The transmittance of the CNF films was measured using UV spectroscopy with a film holder. As

7

shown in Table 1, there is no clear relationship between the drying time and transmittance of the

8

CNF films. Further images of the films and transmittance graphs are reported in the supporting

9

information (S1). The transmittance of the films is an important measurement in regards to

Page 16 of 37

10

imaging and cell staining; if the transmittance of the films is too low, it is difficult to acquire

11

fluorescent microscopy images and reproducibility of measurement might become a concern.

12

Even though there are slight differences in the transmittance of the samples, the transmittance is

13

high enough to allow for staining and microscope imaging following cell culture.

14

The structural characteristics of the CNF substrates were measured using AFM, SEM, water

15

content, PBS uptake, and theromoporometry. It was important to analyze the structure of the CNF

16

substrates to investigate if there is any relationship between structural changes caused by curing

17

at high temperature and the mechanical properties of these films.

18

Atomic force microscopy was performed on the CNF nanopaper to characterize their surface

19

roughness. Figure 3 shows the AFM images of the film surface acquired following drying and

20

curing. The reported roughness values from root-mean-square analysis of AFM images of AC-

21

CNF, 1hC-CNF, and 2hC-CNF (Table 1.) shows that for micro scale roughness there is no

22

significant difference between the roughness. Compared to glass, which was used as a control in

23

the cell aspect ratio and surface area measurements, CNF has higher surface roughness (~4.7

24

nm), while glass is about 0.5 nm as reported by Domke et al. with the same scan size by AFM 46.

ACS Paragon Plus Environment

Page 17 of 37

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

25 26 27

Figure 3. Atomic force microscopy images (10x10 µm) of the surface of the CNF films A) AC-CNF, B) 1hC-CNF, and C) 2hC-CNF taken by tapping mode in air.

28 29

Scanning electron microscopy was performed to determine the thickness and image the porosity

30

of each CNF film as shown in Figure 4. As shown in the figure, the thicknesses of the CNF

31

substrates are similar, but there is a noticeable difference in porosity within the samples, most

32

notably porosity in micrometric scale that can be visualized in the SEM images.

33

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

34

Figure 4. SEM images of the thickness with different magnifications A,B,C) 600x and D,E,F)

35

5000x A,D) AC-CNF, B,E) 1hC-CNF, and C,F) 2hC-CNF with different magnifications and G)

36

measurements of thickness using Image J (mean + SD of 35 membrane section).

Page 18 of 37

37 38

These differences in layer formation could influence the cell adhesion and viability on CNF. The

39

direct nanoscale porosities of the CNF films were measured using thermoporometry, which will

40

be reported later, and the CNF images supports those results with pore size and number

41

decreasing with increased curing time at 150°C. AC-CNF and 1hC-CNF have a more porous

42

structure, while 2hC-CNF is denser with less pores, but still a similar thickness. The direct

43

measurement of this qualitative observation will be discussed later.

44

The water content of the CNF films was determined following conditioning and evaporation of

45

the films at 105°C for 24 hours as outlined in the methods section. This method is commonly

46

used to determine the water content within paper pulp and was adapted for CNF films. As shown

47

in Table 1 curing of CNF films at 150°C for 1 or 2 hours lessens the water content within the film

48

compared to casted CNF (AC-CNF), with maximum loss experienced by 2hC-CNF. This loss of

49

water causes a change in the porosity and the basis weight of the CNF films but have not

50

influenced mechanical properties of the CNF films which are obtained in an aqueous state. PBS

51

absorption of CNF films is dependent on the morphology and structure of the film. Also since the

52

CNF will be tested in liquid conditions and culturing requires the films to be exposed to media,

53

the differences in PBS uptake between the samples can change cell attachment and viability.

54

Since 2hC-CNF has lower porosity compared to AC-CNF and 1hC-CNF the PBS solution cannot

55

easily infiltrate the film and cause swelling. Water content and PBS uptake results are reported in

56

the supporting information (S2)

57

ACS Paragon Plus Environment

Page 19 of 37

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

58 59

Nanoscale porosity of the CNF films in aqueous conditions were determined using

60

thermoporometry. This technique has been employed to characterize cellulosic fibers.47,48 The

61

method was recently improved by Driemeier et al. 45 and is here for the first time used to analyze

62

the nanoscale porosity of CNF films. As expected, with curing at 150°C for either 1 or 2 hours

63

the nanoscale porosity of the films decreased compared to AC-CNF, and further decreased with

64

increased curing time (AC-CNF > 1hC-CNF > 2hC-CNF). The observed loss of porosity is

65

distributed in the measured 1-200 nm pore size range, without special behavior in any specific

66

window of pore size. The maxima of the profiles (FBW read at 200 nm in Figure 5) is also

67

reported in Table 1. The changes in porosity are caused by water evaporation during curing,

68

which causes a decrease in the basis weight of the CNF films. Noteworthy, the changes are

69

irreversible, since the AC-CNF porosity does not recover after 1hC-CNF and 2hC-CNF rehydrate

70

in aqueous conditions. The measured reduction of nanoscale porosity (Figure 5) manifests the

71

cohesive action of drying, which likely impacts other properties, such as CNF film mechanics

72

and PBS uptake, as well as cell growth and viability.

73

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

74

Figure 5. Cumulative distribution of Freezing Bound Water (FBW) (g/g) vs. pore diameter (nm)

75

as measured by thermoporometry for CNF films. For 1hC-CNF and 2hC-CNF, respectively the

76

low and the high branches of the symmetric error bars are omitted for better visualization. With

77

curing the amount of FBW in pores between 1-200 nm decreases.

78 79

In order to attribute the changes in porosity and mechanical properties of the film only to the

80

evaporation of water within the CNF alone, TGA measurements were conducted from 30°C until

81

900°C. Changes in residual mass percentage above 150°C for the samples would indicate a

82

chemical modification of the CNF. As shown in the thermograms, which are reported in the

83

supporting information (S3), there are no differences. This lack of change or shift in the

84

thermograms indicates that the curing of CNF does not produce any chemical change. Therefore,

85

it can be noted that the mechanical properties and porosity changes are only dependent on the

86

curing of the CNF with no chemical modification of the nanocellulose.

87 88

Tuning Mechanical Properties of CNF Substrate

89

Mechanical properties of the CNF films, as well as the viscoelastic properties were determined

90

using tensile tests and DMA with a submersion clamp system. DMA is commonly used to

91

determine if a material used for biomedical applications is suitable, while applying conditions

92

that mimic a physiological environment49. Tensile tests and compression testing were conducted

93

in a reservoir filled with PBS at 37°C to assess the mechanical properties in cell culture

94

conditions. The stress-strain curve and the storage modulus (E’) for the films were measured in

95

hydrated conditions. The stress-strain curves show a significant increase in the Young’s modulus

96

(E) with increase in drying time, and a decrease in elongation at break. These variations in values

ACS Paragon Plus Environment

Page 20 of 37

Page 21 of 37

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

97

show that the samples become more brittle following drying at elevated temperature. The storage

98

modulus was measured using compression from 0.05-20 Hz, which mimics physiological

99

conditions. As shown in Table 1, there was an increase in the storage modulus from 0.05 Hz to 20

100

Hz from AC-CNF to 2hC-CNF with a 56% increase in E’ at 0.05 Hz and 10% increase at 20 Hz

101

for AC-CNF vs. 2hC-CNF. The change in mechanical properties can be attributed to the new

102

nanostructures (described previously) of the CNF substrate since the fiber content and

103

preparation of the films were the same. The only differences occurring in the same trend of

104

increasing mechanical properties is the decrease of porosity with increased drying time, which

105

implies that mechanical performance of CNF films and porosity are related with an R2 value of

106

0.88.

107 108

Figure 6. A) Stress-Strain curves for CNF films tested in phosphate buffer solution at 37°C. The

109

films were either tested as casted (AC-CNF) or after curing (1hC-CNF,2hC-CNF) with (N=3).

110

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

111

The results of mechanical testing in conjunction with CNF film structure characterization (i.e.

112

surface roughness, thickness, porosity) shows that there is a direct relationship between the

113

drying of films for extended periods of time and the material properties. Such results are

114

innovative in regards to CNF substrate material properties, and this tenability has not been

115

studied in details in other scientific paper up to our knowledge. These properties can be tuned

116

depending on the amount of drying time, without any further chemical modification.

117 118

Influence of Tunable CNF substrate on Cell Adhesion and Morphology

119

Since cell adherence and spreading is the first essential step to study the biocompatibility of a

120

material, our next step was to study the adhesion and morphology of D1 MSCs on the tunable

121

CNF substrates. The cell morphology, surface area, and aspect ratio were determined by DAPI-

122

Phalloidin staining of D1 MSCs at 6 h and 24 h following cell seeding. Cells cultured on CNF

123

with different curing conditions displayed different morphologies (Figure 7). Independently the

124

culture time, D1 MSCs seeded on CNF show more elongation and cytoplasmic extension when

125

compared to when seeded on glass (Figure 7 A-H). While cell area at 6 h was not affected by

126

culturing D1 MSCs on any CNF, at 24h only 1hC-CNF and 2hC-CNF showed a higher value

127

when compared to glass (+33%, +28% p