Different Conformations of Surface Cellulose ... - ACS Publications

Sep 27, 2017 - facing the interior had the gauche−trans (gt) conformation. All the other C6-hydroxy groups of the cellulose molecules inside the mic...
1 downloads 0 Views 840KB Size
Subscriber access provided by AUBURN UNIV AUBURN

Article

Different Conformations of Surface Cellulose Molecules in Native Cellulose Microfibrils Revealed by Layer-by-Layer Peeling Ryunosuke Funahashi, Yusuke Okita, Hiromasa Hondo, Mengchen Zhao, Tsuguyuki Saito, and Akira Isogai Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.7b01173 • Publication Date (Web): 27 Sep 2017 Downloaded from http://pubs.acs.org on September 28, 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 26

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

Different Conformations of Surface Cellulose

2

Molecules in Native Cellulose Microfibrils

3

Revealed by Layer-by-Layer Peeling

4 5

Ryunosuke Funahashi, Yusuke Okita, Hiromasa Hondo, Mengchen Zhao, Tsuguyuki Saito, and

6

Akira Isogai*

7 8

Department of Biomaterials Science, Graduate School of Agricultural and Life Sciences, The

9

University of Tokyo, Tokyo 113-8657, Japan

10 11

ABSTRACT: Layer-by-layer peeling of surface molecules of native cellulose microfibrils was

12

performed using a repeated sequential process of 2,2,6,6-tetramethylpiperidine-1-oxyl

13

radical-mediated oxidation followed by hot alkali extraction. Both highly crystalline algal and

14

tunicate celluloses and low-crystalline cotton and wood celluloses were investigated. Initially,

15

the C6-hydroxy groups of the outermost surface molecules of each algal cellulose microfibril

16

facing the exterior had the gauche–gauche (gg) conformation, whereas those facing the interior

17

had the gauche–trans (gt) conformation. All the other C6-hydroxy groups of the cellulose

18

molecules inside the microfibrils contributing to crystalline cellulose I had the trans–gauche

19

(tg) conformation. After surface peeling, the originally 2nd-layer molecules from the

20

microfibril surface became the outermost surface molecules, and the original tg conformation

21

changed to gg and gt conformations. The plant cellulose microfibrils likely had disordered

1 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 2 of 26

22

structures for both the outermost surface and 2nd-layer molecules, as demonstrated using the

23

same layer-by-layer peeling technique.

24 25

KEYWORDS: cellulose microfibril, layer-by-layer peeling, conformation, nuclear magnetic

26

resonance, TEMPO-mediated oxidation

27 28

INTRODUCTION

29

Cellulose, the most abundant extracellular polysaccharide on Earth, is a linear

30

homopolysaccharide consisting of D-glucopyranosyl units linked by β-1,4-glucoside bonds.1

31

Plant cellulose molecules are polymerized by cellulose-synthesizing enzyme complexes in the

32

plasma membrane, directed to the exit channel of the complexes, and crystallized into cellulose

33

microfibrils.2‒9 The cellulose microfibrils are self-assembled and then deposited on primary

34

cell walls as a scaffold to form secondary cell walls, which contribute to the tough physical

35

properties of skeletal tissues.3 Cellulose microfibrils are predominately generated by vascular

36

plants as well as a large number of algae,6 certain bacteria,2,3 and tunicates.5 It has been

37

proposed that plant cellulose microfibrils have disordered surface regions; however, the

38

cross-sectional structures of cellulose microfibrils including the number of cellulose chains in

39

each microfibril remain under debate.10‒15

40

Native cellulose microfibrils have received considerable attention for nanotechnology

41

applications because they are the most abundant reproducible bio-based nanofibers on Earth.

42

Moreover, unlike other inorganic and petroleum-based organic nanomaterials, cellulose

43

microfibrils have unique characteristics such as small widths of 3‒20 nm (depending on the

44

origin of the cellulose), high aspect ratios, high mechanical strength,16 high moduli,17,18 low

2 ACS Paragon Plus Environment

Page 3 of 26

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

45

coefficients of thermal expansion,19 and large surfaces areas. Therefore, many researches have

46

explored methods for efficiently preparing nanosized celluloses (or nanocelluloses) from

47

micrometer-sized plant cellulose fibers and the use of nanocelluloses as fillers and scaffolds for

48

composites and bulk materials such as films, porous materials, and hydrogels.20–25 Some of the

49

resultant materials have exhibited unique mechanical, thermal, optical, catalytic, electric,

50

oxygen-barrier, absorbent, and biological properties, which demonstrate their potential

51

application in high-tech material fields. The nanosized morphologies, characteristics, and

52

functions of the prepared nanocelluloses intrinsically originate from the biosynthesized

53

cellulose microfibrils. The structures of cellulose microfibrils and their molecular-level surface

54

characteristics thus significantly affect the efficient conversion of cellulose fibers into

55

nanocelluloses as well as on their functions and applications.

56

In this study, we investigated the surface structures of highly crystalline cellulose

57

microfibrils from Cladophora sp. and Halocynthia roretzi and low-crystalline microfibrils

58

from cotton lint, cotton linters, and wood cellulose using layer-by-layer peeling of surface

59

molecules. Solid-state

60

conductivity titration, and X-ray diffraction (XRD) analyses of the resultant structures were

61

performed focusing on the conformations of the C6-hydroxy groups of the cellulose

62

microfibrils and their changes during repetitions of the surface peeling process. Furthermore,

63

we determined the local mobility of cellulose molecules and their C6-OH groups in wet and

64

dry states using 13C NMR spectroscopy.

13

C nuclear magnetic resonance (NMR) spectroscopy, electrical

65 66

67

MATERIALS AND METHODS Materials. Marine green alga, Cladophora sp., was collected at the sea of Chikura (Chiba,

3 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 26

68

Japan). A tunicin, Halocynthia roretzi, was obtained at a local fish market. These algal and

69

tunicate celluloses were washed thoroughly with water, soaked in aqueous 0.1 M HCl at RT

70

overnight, and then repeatedly bleached with fresh aqueous 0.3% (w/v) NaClO2 at pH 4–5 and

71

70 °C until the products turned white. These algal and tunicate celluloses were cut into short

72

filaments and small pieces, respectively, using scissors and further purified with 0.1 M HCl

73

and then 4% (w/w) NaOH at room temperature (RT) overnight. Next, 0.5% slurries (200 mL

74

each) of the purified algal and tunicate cellulose in water were mechanically disintegrated at

75

7500 rpm for 10 min using a double-cylinder-type homogenizer (Physcotron NS-56, Microtec

76

Nition, Chiba, Japan). To obtain the cotton lint cellulose, an American cotton ball was

77

purchased from Japan Cotton Promotion Institute (Tokyo, Japan). The cotton lint cellulose was

78

soaked in aqueous 90% (v/v) acetone at RT for 1 day to remove the wax components, followed

79

by rinsing with 90% (v/v) acetone and then thoroughly with water using filtration. The cotton

80

cellulose was cut into short filaments with lengths of ~3 mm, and the wet cellulose (~2 g) was

81

added to a mixture containing NaClO2 (2.28 g) and 0.1 M acetate buffer (200 mL, pH 4.8).

82

Bleaching was performed by stirring the mixture at RT for 3 days, followed by washing

83

thoroughly with water using filtration. A commercial filter pulp (Advantec, Tokyo, Japan) was

84

used as the cotton linters cellulose. A never-dried bleached softwood kraft pulp containing

85

~90% α-cellulose and ~10% mostly hemicelluloses was obtained from Nippon Paper Industry

86

(Tokyo, Japan). All the reagents and solvents were laboratory grade (Wako Pure Chemicals,

87

Japan) and used as received.

88

TEMPO-Mediated Oxidation. The wet cellulose sample with a weight corresponding to a

89

dry weight of 2 g was suspended in water (200 mL) containing TEMPO (0.032 g) and NaBr

90

(0.2 g).26,27 The oxidation was started by adding 1.8 M NaClO solution to the cellulose

91

suspensions: 10 mmol/g NaClO for the algal, tunicate, and cotton celluloses and 3.8 mmol/g 4 ACS Paragon Plus Environment

Page 5 of 26

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

92

NaClO for the wood cellulose. The suspension was maintained at pH 10 by continuous

93

addition of 0.5 M NaOH using a pH stat (AUT-701, DKK-TOA, Tokyo, Japan) for 4 h for the

94

algal, tunicate, and cotton celluloses and for 40 min for the wood cellulose. The oxidized

95

celluloses were thoroughly washed with water using filtration and stored at 4 °C without

96

drying.

97

TEMPO-Oxidized Cellulose Nanofibrils. The 0.1% (w/v) TEMPO-oxidized algal, tunicate,

98

cotton, and wood cellulose/water slurries were mechanically disintegrated at 7500 rpm for 2 min

99

using the double-cylinder-type homogenizer. The obtained TEMPO-oxidized cellulose

100

nanofibril/water dispersions were then sonicated for 6‒12 min using an ultrasonic homogenizer

101

with a 7-mm probe tip diameter at 19.5 kHz and 300-W output power (US-300T, Nihon Seiki,

102

Tokyo, Japan). The unfibrillated fraction, if present in the dispersion, was removed by

103

centrifugation at 12 000×g for 10 min.

104

Layer-By-Layer Surface Peeling. A solution of 2 M aqueous NaOH (200 mL) was slowly

105

poured into the 0.2% TEMPO-oxidized cellulose suspension (200 mL) containing NaBH4

106

(0.01 g). The mixture was kept at 105 °C for 12 h after bubbling N2 gas into the mixture for 10

107

min. After being washed thoroughly with water using centrifugation, the water-insoluble

108

residue was oxidized again using the same TEMPO/NaBr/NaClO system in water at pH 10

109

described above, and the obtained TEMPO-oxidized cellulose was then extracted with 1 M

110

NaOH under the same conditions described above.

111

Analyses. The carboxylate contents of the oxidized celluloses were determined using the

112

electrical conductivity titration method.28 AFM images of the TEMPO-oxidized cellulose

113

nanofibrils were captured using an atomic force microscope (NanoScope IIIa, Veeco

114

Instruments, Inc., USA) with a Bruker MPP-11100-10 tip operating in tapping mode. The largest

115

height value measured along each isolated nanofibril was taken as the true height value, as the 5 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 6 of 26

116

tapping step (512×512 pixel images for 5-µm-wide squares) was large compared with the narrow

117

(~10 nm wide) nanofibrils. More than 60 nanofibrils were measured for each sample. The

118

freeze-dried samples (~0.1 g each) were pressed at approximately 750 MPa for 1 min to make

119

pellets. XRD patterns were recorded for the pellets for 2θ diffraction angles between 10° and 30°

120

in reflection mode using a diffractometer (RINT 2000, Rigaku, Tokyo, Japan) with Ni-filtered

121

Cu Kα radiation (λ = 0.1542 nm) at 40 kV and 40 mA. The crystal widths of cellulose I were

122

calculated from the full widths at half maximums of the (2 0 0) diffraction peaks of the tunicate,

123

cotton and wood celluloses using Scherrer’s equation.29 Two peaks centered at approximately

124

14.8° and 16.8° in the XRD patterns of the algal celluloses, which correspond to d spacings of

125

0.60‒0.61 and 0.53‒0.54 nm, respectively, were deconvoluted via curve-fitting using a

126

pseudo-Voigt function.30 The water contents of the never-dried and re-wetted samples inside an NMR rotor were

127 128

approximately

129

Cross-polarization/magic angle sample spinning (CP/MAS)

130

performed using an NMR spectrometer (JNM-ECAII 500, JEOL, Japan) operating at 125.77

131

MHz for

132

performed at 298 K under the following conditions: sample spinning rate of 6 kHz, proton 90°

133

pulse time of 2.5 µs, and relaxation delay of 5 s. The CP transfer was achieved using a ramped

134

amplitude sequence (RAMP/CP) for a CP contact time of 2 ms and an amplitude graduation of

135

7%. Adamantane was used as an external standard for the ppm calculation. Each CP/MAS

136

spectrum was deconvoluted into Gaussian and Lorentzian components (i.e., Voigt components)

137

at 60‒62, 62.5‒64.5, and 65.6‒66.6 ppm assigned to C6-OH conformations of gg, gt, and tg,

138

respectively, using FeakFit version 4.12 (Seasolve Software Inc., San Jose, CA, USA).31,32 The

139

13

13

50‒60%;

the

dried

samples

were

prepared 13

after

freeze-drying.

C NMR measurements were

C with a 3.2-mm HXMAS probe and a ZrO2 rotor. All the measurements were

C T1-measurement was performed using the pulse sequence developed by Torchia,33 with a 6 ACS Paragon Plus Environment

Page 7 of 26

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

140

relaxation delay of 10 s and eight τ values between 0.1 and 60 s for the cotton and wood

141

celluloses. The

142

between 0.1 and 400 s. The 13C T1 values were calculated based on the time-dependent signal

143

intensities. The evolution of the signal intensity with τ was modeled using weighted linear

144

least-squares fitting.

13

C T1 measurements of algal cellulose were performed with eight τ values

145 146

RESULTS AND DISCUSSION

147

TEMPO-Mediated Oxidation and Alkali Treatment of Algal Cellulose. Almost all the

148

C6-hydroxy groups exposed on crystalline cellulose microfibril surfaces in native celluloses

149

are position-selectively oxidized to Na C6-carboxylate groups by TEMPO-mediated oxidation

150

under suitable conditions.22,27 As a result, one of every two glucosyl units on the cellulose

151

microfibril surfaces are mostly converted into sodium glucuronosyl units by the

152

TEMPO-mediated

153

C6-carboxylate groups on the TEMPO-oxidized cellulose microfibril surfaces could thus be

154

removed by alkali extraction. Native celluloses are stable without swelling and most glycoside

155

bonds in cellulose molecules are not significantly cleaved with treatment using 1 M NaOH at

156

105°C.38 Thus, the treatment with 1 M NaOH at 105°C for 12 h was used to remove as many

157

of the C6-carboxylate-group-containing cellulose molecules as possible. This treatment was

158

much harsher than that previously used to obtain glucose/glucuronic acid alternating

159

co-polysaccharides from TEMPO-oxidized native celluloses by surface peeling with 10%

160

(w/w) NaOH at RT for 30 min.37 As a result, new cellulose microfibrils with one surface layer

161

less than the original microfibrils were expected to be formed by the surface peeling (Figure 1).

oxidation.22,27,34‒37

The

surface

cellulose

molecules

containing

162

7 ACS Paragon Plus Environment

Biomacromolecules

163 164 165 166 167 168

Figure 1 Schematic model of layer-by-layer peeling of cellulose microfibril surface by

169

TEMPO-mediated oxidation and subsequent alkali extraction.

170 0.6

171 172 173 174 175 176

Carboxylate content (mmol/g)

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 26

0.5 0.4 0.3 0.2 0.1 0.0

Or i gi

na l

1s to

xi d i

1s ta ze d

2n 2n 3rd 3rd do da ox alk lka lka xi d i di ali li -e li -e ize ze -ex xtra d d xtra tra cte cte cte d d d

177

Figure 2 Carboxylate contents of algal cellulose samples prepared by repetitions of

178

TEMPO-mediated oxidation and subsequent hot alkali extraction.

179 180

The carboxylate content of the 1st TEMPO-oxidized algal cellulose was 0.51 mmol/g, which

181

is consistent with previously reported results.36 This carboxylate content decreased to 0.06

182

mmol/g after the 1st hot alkali treatment. When this 1st TEMPO-oxidized and alkali-treated

183

sample was re-oxidized using the TEMPO/NaBr/NaClO system, the resulting 2nd

184

TEMPO-oxidized sample had a carboxylate content of 0.53 mmol/g. This value decreased to

185

0.06 mmol/g after the 2nd hot alkali treatment. The carboxylate contents of the original,

186

TEMPO-oxidized, and TEMPO-oxidized/alkali-treated algal cellulose samples up to three

187

repetitions are shown in Figure 2. Although complete removal of carboxylate groups in 8 ACS Paragon Plus Environment

Page 9 of 26

188

TEMPO-oxidized algal celluloses could not be achieved in each hot alkali treatment,

189

approximately 90% of the carboxylate groups present in each TEMPO-oxidized algal cellulose

190

sample were removed by the hot alkali treatment.

191

XRD and Atomic Force Microscopy Analyses of TEMPO-Oxidized/Alkali-Treated

192

Algal Celluloses. The cellulose microfibril widths of the original, TEMPO-oxidized, and

193

alkali-treated algal cellulose samples after the 1st, 2nd, and 3rd treatments shown in Figure 2

194

were examined using XRD and AFM. The XRD patterns of the algal cellulose samples are

195

presented in Figure S1 (see Supporting Information). The changes in the crystal sizes

196

corresponding to the diffraction peaks at 2θ = 14.8° and 16.8° for the algal cellulose samples

197

are shown in Figure 3. The cross sections of the algal cellulose microfibrils were assumed to be

198

rectangular or almost square,39–44 and the crystal sizes were calculated from the XRD patterns

199

based on the microfibril model shown in Figure S2 using Scherrer’s equation.29

200 20

201 202 203 204 205 206

Cellulose I crystal width (nm)

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

15

10

5 Crystal width at 14.8 ° (1 1 0) Crystal width at 16.8 ° (1 -1 0)

0

Or i gi

na l

1s

to

xi d i

1s ta ze d

2n 2n 3rd 3rd da do alk ox lka xi d lka id i alili -e li -e i ze z e ex xtr d d xtra tra ac cte cte ted d d

207

Figure 3 Crystal widths of (1 1 0) and (1 ‒1 0) planes of cellulose I of algal cellulose samples

208

prepared by repetitions of TEMPO-mediated oxidation and subsequent hot alkali extraction.

209 210

The original crystal sizes of the (1 1 0) and (1 ‒1 0) planes were unchanged after the 1st

211

TEMPO-mediated oxidation; however, these sizes decreased by ~1 nm after the 1st hot alkali 9 ACS Paragon Plus Environment

Biomacromolecules

212

treatment. This value was unchanged after the 2nd TEMPO-mediated oxidation but decreased

213

by ~1 nm after the 2nd hot alkali treatment. Similar results were observed for the 3rd

214

TEMPO-oxidized and alkali-extracted algal cellulose samples (Figure 3). These XRD results

215

indicate that the crystal size of the algal cellulose decreased by ~1 nm after each round of

216

TEMPO-mediated oxidation and subsequent hot alkali extraction.

217 218 219 220 221 222

Cellulose nanofibril width measured from AFM height image (nm)

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 26

20

15

10

5

0

Or igin al

1s 1s 2n 2n 3rd 3rd to ta do da ox alk xid lka lka xid idiz aliize li-e li-e zie ed ex d xtr d x tra tra ac cte c ted ted d

223

Figure 4 TEMPO-oxidized cellulose nanofibril widths measured from AFM height images of

224

algal cellulose samples prepared by repetitions of TEMPO-mediated oxidation and subsequent

225

hot alkali extraction.

226 227

The TEMPO-oxidized algal celluloses prepared by the 1st, 2nd, and 3rd treatments shown in

228

Figures 2 and 3 were converted into individual nanofibrils dispersed in water using mechanical

229

disintegration.22,26,27 AFM examination was performed to measure the average heights of

230

individual TEMPO-oxidized cellulose nanofibrils (Figure S3). The average microfibril size of

231

the 1st TEMPO-oxidized algal cellulose nanofibrils was 14.2 ± 2.2 nm (Figure 4). This size

232

decreased to 13.3 ± 2.2 nm after the 1st hot alkali treatment and the 2nd TEMPO-mediated

233

oxidation. This size further decreased to 12.2 ± 3.0 nm with the 2nd hot alkali treatment and

234

subsequent TEMPO-mediated oxidation. These AFM observations of the TEMPO-oxidized

235

algal cellulose nanofibrils indicate that the microfibril width decreased by ~1 nm with one

10 ACS Paragon Plus Environment

Page 11 of 26

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

236

sequential treatment of TEMPO-mediated oxidation/hot-alkali extraction, as also observed in

237

the XRD results presented in Figure 3.

238

The results in Figures 2‒4 show that the layer-by-layer peeling of the surface molecules of

239

algal cellulose microfibrils was successful with repetitions of the TEMPO-mediated oxidation

240

followed by hot alkaline extraction. This peeling was possible because the thicknesses of each

241

layer of a cellulose I crystallite corresponding to the (1 1 0) and (1 ‒1 0) planes are 0.53 and

242

0.61 nm, respectively, or ~0.5 nm (Figure S2).1,43 The alkali-soluble fractions obtained from

243

TEMPO-oxidized algal celluloses at each stage were highly degraded and depolymerized under

244

the alkaline conditions used.45

245

Similar results were obtained for highly crystalline tunicate cellulose, which has pure

246

cellulose Iβ and a parallelogram cross-section of microfibrils,46‒50 in layer-by-layer peeling

247

experiments (Figures S4–S6).

248

Solid-State 13C NMR Spectroscopy Analysis of TEMPO-Oxidized/Alkali-Treated Algal

249

Celluloses. The hydroxymethyl or C6-OH groups can have three conformations in cellulose

250

microfibrils: trans–gauche (tg), gauche–trans (gt), and gauche–gauche (gg) appearing at 60‒62,

251

62.5‒64.5, and 65.6‒66.6 ppm, respectively, in solid-state 13C NMR spectra.31,32,48. The C6-OH

252

signals of the original and 1st TEMPO-oxidized algal celluloses are presented in Figures 5A

253

and 5B, respectively. The signal for the C6-OH groups with the gg conformation appeared at

254

61.6‒61.7 ppm for the original algal cellulose, and its intensity decreased after the

255

TEMPO-mediated oxidation. The C6-OH groups present on the cellulose microfibril surfaces

256

were position-selectively converted into Na C6-carboxylate groups by the TEMPO-mediated

257

oxidation.22,26,27,34,35 Thus, the C6-OH groups present on the algal cellulose microfibril surfaces

258

likely had the gg conformation, which is consistent with a previously reported hypothesis.51

259 11 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 12 of 26

260 261 262 263 264 265 266 267 268 269 270 271 272

Figure 5 Solid-state 13C NMR spectra of C6 signals of the original algal cellulose (A) and the

273

1st TEMPO-oxidized algal cellulose (B), with their deconvoluted patterns to tg, gt, and gg

274

conformations. The molar ratios of the tg, gt, and gg conformations of C6-OH groups of algal

275

cellulose samples prepared by repetitions of TEMPO-mediated oxidation and subsequent hot

276

alkali extraction (C) from solid-state 13C NMR spectra.

277 278

The carboxylate content of the 1st TEMPO-oxidized algal cellulose was 0.51 mmol/g

279

(Figure 2), which corresponds to 0.084 mol/mol of repeating units (i.e., glucosyl and

280

glucuronosyl units) according to a previously reported calculation (see Eq. S1 in Supporting

281

Information).36 The 8.4% glucosyl units in the original algal cellulose were converted into

282

C6-carboxylate units by the 1st TEMPO-mediated oxidation. As shown in Figure 5C, the

283

fraction of gg conformation in the original algal cellulose measured by solid-state

284

spectroscopy analysis was 6.9%, which is close to 8.4%. These results indicate that one of

13

C NMR

12 ACS Paragon Plus Environment

Page 13 of 26

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

285

every two C6-OH groups in the molecules present on each algal cellulose microfibril surface

286

facing the exterior was likely to have the gg conformation.52‒54

287

Note that the gg:gt conformation ratios in the original algal celluloses as well as the 1st, 2nd,

288

and 3rd alkali-extracted algal cellulose samples were almost 1:1 (Figure 5C). This result

289

indicates that one of every two C6-OH groups in the molecules on each algal cellulose

290

microfibril surface facing toward the interior of the microfibril had the gt conformation (Figure

291

6). The glucosyl units with gg and gt conformations of C6-OH groups were, therefore,

292

alternatingly linked to each other in each cellulose molecule present on the algal microfibril

293

surface.37 Based on the cross-sectional model of the algal cellulose microfibril,1,36,39–44 16% of

294

the glucosyl units were present on the microfibril surfaces, and 84% were present inside the

295

microfibrils. Figure 5C shows that 86% of the glucosyl units in the 1st TEMPO-oxidized algal

296

cellulose had the tg conformation, which is consistent with the XRD data.

297 298 299 300 301 302 303 304

Figure 6 Schematic model of C6-OH conformations in the 1st, 2nd, and 3rd layers from the

305

algal cellulose microfibril surface..

306 307

The same interpretations were made for the 2nd and 3rd TEMPO-oxidized algal celluloses

308

and 2nd and 3rd alkali-extracted algal celluloses based on the layer-by-layer peeling model of 13 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 14 of 26

309

the algal cellulose microfibrils by the TEMPO-oxidation/alkali-extraction process (Figures 5C,

310

S7, and S8). Thus, the cellulose molecules inside the algal cellulose microfibril surfaces always

311

had the tg conformation as cellulose I crystallites.1 A schematic model of the layer-by-layer

312

surface peeling of the algal cellulose microfibrils is presented in Figure 7.

313 314 315 316 317 318 319

Figure 7 Schematic model of the layer-by-layer peeling of algal cellulose microfibril surface

320

by repetitions of TEMPO-mediated oxidation and subsequent hot alkali extraction.

321 322

When the cellulose molecules originally present in the 2nd or 3rd layer from the microfibril

323

surface were exposed to the outermost surface by the layer-by-layer peeling, the gg and gt

324

conformations again appeared at the same 1:1 ratio (Figure 7). The signal patterns of C1

325

carbon at ~105 ppm in the solid-state

326

indicate that the ratio of cellulose Iα/Iβ was unchanged during the repetitions of the

327

TEMPO-oxidation/alkali extraction treatment (Figures S 7 and S8).32,55,56 Moreover, the gg, gt,

328

and tg conformation ratios for the original and alkali-extracted samples shown in Figure 5C

329

were almost unchanged between the wet and dry states (Figure S9).

330

13

C NMR spectra for these algal cellulose samples

Spin-Lattice Relaxation Times of Algal Cellulose Carbons. The

13

C T1 relaxation times

331

were measured using a pulse-sequence developed by Torchia.33 13C atoms with more freedom

332

of movement have shorter 13C T1 values, whereas those in rigidly ordered domains have longer 14 ACS Paragon Plus Environment

Page 15 of 26

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

333

13

C T1 values.57

13

334

including disordered regions. For the highly crystalline algal cellulose microfibril, the mobile

335

components should originate from the outermost microfibril surfaces, whereas the rigid

336

components should originate from inside the crystalline regions. As described in the previous

337

section, the C6-OH groups with the gg and gt conformations were located on the outermost

338

microfibril surfaces, whereas those with the tg conformation were located inside the crystalline

339

microfibrils (Figures 6 and 7).

C atoms with short T1 values are found on cellulose microfibril surfaces

340 341

Table 1. T1 Relaxation Times (ms) of C4 and C6 Carbons of Algal Cellulose.

C4

Never-dried

Freeze-dried

35

8.0

Freeze-dried and then re-wetted 25

1700

860

1600

gg conformation

0.4

2.0

0.4

tg conformation

940

430

1100

Surface Inside

C6 342 343

The T1 values of the C4 and C6 carbons in the original algal cellulose in the never-dried,

344

freeze-dried, and re-wetted states are listed in Table 1. These 13C T1 values clearly demonstrate

345

that the C4 and C6 carbons at the outermost surfaces of the microfibrils had much shorter T1

346

values than those inside the microfibril or crystalline regions for all three dry and wet states.

347

This result is reasonable because of the different molecular environments between the

348

microfibril surfaces and inside the crystallites. Moreover, the local mobility of C4 and C6

349

carbons (except for the C6 carbons with the gg conformation) was much more highly restricted

350

in the wet state than in the dry state. This difference in mobility can possibly be explained

351

using the following hypothesis: the hydration of algal cellulose microfibril surfaces or

15 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 16 of 26

352

hydrogen bond formation between water molecules and C6-OH groups with the gg

353

conformation on the microfibril surfaces improve the rigidity of C4 carbons of both microfibril

354

surfaces and interiors and of C6 carbons inside the microfibrils. Because the C6-OH groups

355

with the gg conformation on the microfibril surface could form hydrogen bonds with water

356

molecules in the wet states, their

357

states. In contrast, there may not be opportunities for the C4 carbons on the microfibril surfaces

358

to form hydrogen bonds with water molecules even in the wet states, resulting in the surface

359

C4 carbons having higher T1 values in the wet states than in the dry state.

13

C T1 values were higher in the dry state than in the wet

360

Layer-by-Layer Peeling of Cotton and Wood Celluloses. It has been proposed that plant

361

cellulose microfibrils have more disordered structures on their surfaces than highly crystalline

362

algal cellulose.7‒12 The layer-by-layer peeling technique described above was thus applied to

363

cotton and wood celluloses. The carboxylate contents of the 1st and 2nd TEMPO-oxidized

364

cotton and wood celluloses and those of the 1st and 2nd hot alkali extractions are shown in

365

Figure 8A. Significant amounts of carboxylate groups were formed in the TEMPO-oxidized

366

celluloses, and most of them were removed by the hot alkali extraction, as for the algal and

367

tunicate celluloses. These results indicate that the layer-by-layer peeling of the

368

TEMPO-oxidized cellulose molecules in the plant cellulose microfibrils is likely to be

369

achieved by the repetition of the TEMPO-mediated oxidation and subsequent hot alkali

370

extraction process.

371

The crystal sizes of the original cotton and wood celluloses did not, however, decrease by ~1

372

nm after each hot alkali extraction, in contrast to the results for the algal and tunicate celluloses

373

(Figure 8B). These findings indicate that the plant cellulose microfibrils have disordered

374

structures not only in the 1st layer but also in part of the 2nd layer from the microfibril surface.

375

The solid-state

13

C NMR spectra of C6 carbons of these plant celluloses after 16 ACS Paragon Plus Environment

Page 17 of 26

376

TEMPO-mediated oxidation and hot alkali extraction were too broad in both the dry and wet

377

states for accurate signal deconvolution to separate the tg, gt, and gg conformations. Further

378

studies are thus needed to investigate the layer-by-layer peeling of these plant celluloses.

379 380 381 382 383

Carboxylate content (mmol/g)

2.0

384

A

1st oxidized 1st alkali-extracted 2nd oxidized 2nd alkali-extracted

1.5

1.0

0.5

0.0 Wood cellulose

385

B

386 387 388 389 390

Cotton lint cellulose

Cotton linters cellulose

20

Cellulose I crystal width (nm)

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

Cotton linters Cotton lint Wood cellulose

15

10

5

0

Or

igi

na

l

1s to

391

xid iz

1s ed

ta

lka

2n li- e

do

xtr ac t ed

2n xid i

ze d

da

lka

li- e

xtr a

cte

d

392

Figure 8 Carboxylate contents of wood, cotton lint, and cotton linters cellulose samples

393

prepared by repetitions of TEMPO-mediated oxidation and subsequent hot-alkali extraction

394

(A) and corresponding changes in crystal width of the (2 0 0) plane of cellulose I (B).

395 396

CONCLUSION

397

The layer-by-layer peeling and solid-state 13C NMR analysis of algal cellulose revealed that the

398

C6-OH groups of surface cellulose molecules facing both outside and inside the microfibrils

399

have alternating gg and gt conformations, respectively, even after layer-by-layer peeling. In 17 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 18 of 26

400

contrast, the C6-OH groups inside the microfibrils forming cellulose I crystallites always had

401

the tg conformation. The weight ratios of the three C6-OH conformations in the original algal

402

cellulose and those of the TEMPO-oxidized/hot alkali extracted samples remained almost

403

unchanged between the wet and dry states. The local mobility of entire cellulose molecules

404

except for C6 carbons with the gg conformation, evaluated by the

405

restricted in the wet state. Thus, the conformations of the C6-OH groups and their T1 values in

406

cellulose molecules were different for the microfibril surface and interior. When the

407

layer-by-layer peeling technique was applied to low-crystalline plant celluloses, no clear

408

decrease in the cellulose I crystal width was observed, indicating that the molecules in the 1st

409

and 2nd layers from the surface of these plant cellulose microfibrils have disordered structures.

410

The surface-peeling technique developed in this study can thus be used to perform structural

411

analyses of native cellulose microfibrils and to prepare new nanocelluloses with specific

412

widths.

13

C T1 values, was highly

413 414

ASSOCIATED CONTENT

415

Supporting Information

416

Calculation model for cellulose crystal sizes; XRD patterns, AFM images, and solid-state 13C

417

NMR spectra of layer-by-layer surface peeled algal celluloses; changes in carboxylate content

418

and microfibril width of layer-by-layer-surface peeled tunicate cellulose; molar ratios of tg, gt,

419

and gg conformations of C6-OH groups of surface-peeled algal celluloses in dry and wet states.

420 421

AUTHOR INFORMATION

422

Corresponding Author

423

*Tel: +81 3 5841 5538. E-mail: [email protected] 18 ACS Paragon Plus Environment

Page 19 of 26

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

424 425

Notes

426

The authors declare no competing financial interest.

427 428

ACKNOWLEDGMENTS

429

This research was supported by Core Research for Evolutional Science and Technology

430

(CREST, Grant number JPMJCR13B2) of the Japan Science and Technology Agency (JST).

431 432

REFERENCES

433

(1)

434

Bonding System in Cellulose Iα from Synchrotron X-Ray and Neutron Fiber Diffraction. J. Am.

435

Chem. Soc. 2003, 125, 14300‒14306.

436

(2)

437

and Orienting Complexes in Association with the Plasma Membrane. Proc. Natl. Acad. Sci.

438

USA 1976, 73, 143‒147.

439

(3)

440

Acetobacter xylinum: Visualization of the Site of Synthesis and Direct Measurement of the in

441

vivo Process. Proc. Natl. Acad. Sci. USA 1976, 73, 4565‒4569.

442

(4)

Delmer, D. P.; Amor, Y. Cellulose Biosynthesis. Plant Cell 1995, 7, 987‒1000.

443

(5)

Kimura, S.; Itoh, T. New Cellulose Synthesizing Complexes (Terminal Complexes)

444

Involved in Animal Cellulose Biosynthesis in the Tunicate Metandrocarpa uedai. Protoplasma

445

1996, 194, 151‒163.

446

(6)

447

1996, 33, 1345‒1373.

Nishiyama, Y.; Sugiyama, J.; Chanzy, H.; Langan, P. Crystal Structure and Hydrogen

Brown Jr., R. M.; Montezinos, D. Cellulose Microfibrils: Visualization of Biosynthetic

Brown Jr., R. M.; Willison, J. H.; Richardson, C. L. Cellulose Biosynthesis in

Brown, R. M., Jr. The Biosynthesis of Cellulose. J. Macromol. Sci.-Pure Appl. Chem. A

19 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 20 of 26

448

(7)

Koyama, M.; Helbert., W.; Imai, T.; Sugiyama, J.; Henrissat, B. Parallel-up Structure

449

Evidences the Molecular Directionality During Biosynthesis of Bacterial Cellulose. Proc. Natl.

450

Acad. Sci. USA 1997, 94, 9091‒9095.

451

(8)

452

Curr. Opin. Plant Biol. 2004, 7, 651‒660.

453

(9)

454

and Membrane Translocation. Nature 2013, 493, 181‒186.

455

(10) Paredez, A. R.; Somerville, C. R.; Ehrhardt, D. W. Visualization of Cellulose Synthase

456

Demonstrates Functional Association with Microtubules. Science 2006, 312, 1491‒1495.

457

(11) Ding, S.-Y.; Himmel, M. E. The Maize Primary Cell Wall Microfibril: A New Model

458

Derived from Direct Visualization. J. Agr. Food Chem. 2006, 54, 597–606.

459

(12) Nixon, B. T.; Mansouri, K.; Singh, A.; Du, J.; Davis, J.; Lee, J.-G.; Slabaugh, E.;

460

Vandavasi, V. G.; O’Neill, H.; Roberts, E. M.; Roberts, A. W.; Yingling, Y. G.; Haigler, C. H.

461

Comparative Structural and Computational Analysis Supports Eighteen Cellulose Synthases in

462

the Plant Cellulose Synthesis Complex. Sci. Rep. 2016, 6, 28696.

463

(13) Jarvis, M. C. Cellulose Biosynthesis: Counting the Chains. Plant Physiol. 2013, 163,

464

1485–1486.

465

(14) Zhang, T.; Zheng, Y.; Cosgrove, D. J. Spatial Organization of Cellulose Microfibrils and

466

Matrix Polysaccharides in Primary Plant Cell Walls as Imaged by Multichannel Atomic Force

467

Microscopy. Plant J. 2016, 85, 179‒192.

468

(15) Silveira, R. L.; Stoyanov, S. R.; Kovalenko, A.; Skaf, M. S. Cellulose Aggregation under

469

Hydrothermal Pretreatment Conditions. Biomacromolecules 2016, 17, 2582−2590.

470

(16) Saito, T.; Kuramae, R.; Wohlert, J.; Berglund, L. A.; Isogai, A. An Ultrastrong

471

Nanofibrillar Biomaterial: the Strength of Single Cellulose Nanofibrils Revealed via

Wasteneys, G. O. Progress in Understanding the Role of Microtubules in Plant Cells.

Morgan, J. L.; Strumillo, J.; Zimmer, J. Crystallographic Snapshot of Cellulose Synthesis

20 ACS Paragon Plus Environment

Page 21 of 26

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

472

Sonication-Induced Fragmentation. Biomacromolecules 2013, 14, 248‒253.

473

(17) Sturcová, A.; Davies, G. R.; Eichhorn, S. J. The Elastic Modulus and Stress-Transfer

474

Properties of Tunicate Cellulose Whiskers. Biomacromolecules 2005, 6, 1055‒1061.

475

(18) Iwamoto, S.; Kai, W.; Isogai, A.; Iwata, T. Elastic Modulus of Single Cellulose

476

Microfibrils from Tunicate Measured by Atomic Force Microscopy. Biomacromolecules 2009,

477

10, 2571‒2576.

478

(19) Hori, R.; Wada, M. The Thermal Expansion of Wood Cellulose Crystals. Cellulose 2005,

479

12, 479−484.

480

(20) Habibi, Y.; Lucian, L.; Rojas, O. J. Cellulose Nanocrystals: Chemistry, Self-Assembly,

481

and Applications. Chem. Rev. 2010, 110, 3479–3500.

482

(21) Siró, I.; Plackett, D. Microfibrillated Cellulose and Bew Nanocomposite Materials: a

483

review. Cellulose 2010, 17, 459–494.

484

(22) Isogai, A.; Saito, T.; Fukuzumi, H. TEMPO-Oxidized Cellulose Nanofibers. Nanoscale

485

2011, 3, 71–85.

486

(23) Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A.

487

Nanocelluloses: A New Family of Nature-Based Materials. Angew. Chem. 2011, 50, 5438–

488

5466.

489

(24) Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Yungblood, J. Cellulose Nanomaterials

490

Review: Structure, Properties and Nanocomposites. Chem. Soc. Rev. 2011, 40, 3941–3994.

491

(25) Isogai, A. Wood Nanocelluloses: Fundamentals and Applications as New Bio-Based

492

Nanomaterials. J. Wood Sci. 2013, 59, 449–459.

493

(26) Saito, T.; Kimura, S.; Nishiyama, Y.; Isogai, A. Cellulose Nanofibers Prepared by

494

TEMPO-Mediated Oxidation of Native Cellulose. Biomacromolecules 2007, 8, 2485–2491.

495

(27) Shinoda, R.; Saito, T.; Okita, Y.; Isogai, A. Relationship between Length and Degree of 21 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 22 of 26

496

Polymerization of TEMPO-Oxidized Cellulose Nanofibrils. Biomacromolecules 2012, 13,

497

842–849.

498

(28) Saito, T.; Isogai, A. TEMPO-Mediated Oxidation of Native Cellulose. The Effect of

499

Oxidation Conditions on Chemical and Crystal Structures of the Water-Insoluble Fractions.

500

Biomacromolecules 2004, 5, 1983–1989.

501

(29) Alexander, L. E. “X-Ray Diffraction Methods in Polymer Science”, ed. Robert, E.,

502

Krieger Publishing, New York, 1979, pp 423‒424.

503

(30) Wada, M.; Okano, T.; Sugiyama, J. Synchrotron-Radiated X-Ray and Neutron

504

Diffraction Study of Native Cellulose. Cellulose 1997, 4, 221–232.

505

(31) Horii, F.; Hirai, A.; Kitamaru, R. Solid-State

506

Oligosaccharides and Cellulose - Conformation of CH2OH Group about the Exo-Cyclic C-C

507

Bond. Polym. Bull. 1983, 10, 357‒361.

508

(32) Larsson, P. T.; Westlund, P. O. Line Shapes in CP/MAS 13C NMR Spectra of Cellulose I.

509

Spectrochim. Acta A 2005, 62, 539–546.

510

(33) Torchia, D. A. The Measurement of Proton-Enhanced Carbon-13 T1 Values by a Method

511

Which Suppresses Artifacts. J. Magn. Reson. 1978, 30, 613‒616.

512

(34) Montanari, S.; Roumani, M.; Heux, L.; Vignon, M. R. Topochemistry of Carboxylate

513

cellulose Nanocrystals Resulting from TEMPO-Mediated Oxidation. Macromolecules 2005, 38,

514

1665‒1671.

515

(35) Saito, T.; Hirota, M.; Tamura, N.; Kimura, S.; Fukuzumi, H.; Heux, L.; Isogai, A.

516

Individualization of Nano-Sized Plant Cellulose Fibers by Direct Surface Carboxylation Using

517

TEMPO Catalyst under Neutral Conditions. Biomacromolecules 2009, 10, 1992–1996.

518

(36) Okita, Y.; Saito, T.; Isogai, A. Entire Surface Oxidation of Various Cellulose

519

Microfibrils by TEMPO-Mediated Oxidation. Biomacromolecules 2010, 11, 696‒1700.

13

C-NMR Study of Conformations of

22 ACS Paragon Plus Environment

Page 23 of 26

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

520

(37) Hirota, M.; Furihata, K.; Saito, T.; Kawada, T.; Isogai, A. Glucose/Glucuronic Acid

521

Alternating Co-Polysaccharides Prepared from TEMPO-Oxidized Native Celluloses by

522

Surface Peeling. Angew. Chem. Int. Ed. 2010, 49, 7670‒7673.

523

(38) Krässig, H. A. “Polymer monographs. In: Cellulose: Structure, Accessibility and

524

Reactivity”, ed. Huglin, M. B., Gordon and Breach Science Publishers, Amsterdam, 1993, Vol

525

11, pp 264.

526

(39) Revol, J.-F. On the Cross-Sectional Shape of Cellulose Crystallites in Valonia ventricosa.

527

Carbohydr. Polym. 1982, 2, 123‒134.

528

(40) Sugiyama, J.; Harada, H.; Fujiyoshi, Y.; Uyeda, N. Lattice Images from Ultrathin

529

Sections of Cellulose Microfibrils in the Cell Wall of Valonia macrophysa Ktüz. Planta 1985,

530

166, 161‒168.

531

(41) Hanley, S. J.; Giasson, J.; Revol, J.-F.; Gray, D. G. Atomic Force Microscopy of

532

Cellulose Microfibrils: Comparison with Transmission Electron Microscopy. Polymer 1992,

533

33, 4639‒4642.

534

(42) Imai, T.; Putaux, J. L.; Sugiyama, J. Geometric Phase Analysis of Lattice Images from

535

Algal Cellulose Microfibrils. Polymer 2003, 44, 1871‒1879.

536

(43) Nishiyama, Y. Structure and Properties of the Cellulose Microfibril. J. Wood. Sci. 2009,

537

55, 241‒249.

538

(44) Nishiyama, Y.; Wada, M.; Hanson, B. L.; Langan, P. Time-Resolved X-Ray Diffraction

539

Microprobe Studies of the Conversion of Cellulose I to Ethylenediamine-Cellulose I. Cellulose

540

2010, 17, 735–745.

541

(45) Fujisawa, S.; Isogai, T.; Isogai, A. Temperature and pH Stability of Cellouronic Acid.

542

Cellulose 2010, 17, 607–615.

543

(46) Sugiyama, J.; Persson, J.; Chanzy, H. Combined Infrared and Electron Diffraction Study 23 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 24 of 26

544

of the Polymorphism of Native Celluloses. Macromolecules 1991, 24, 2461–2466.

545

(47) Wada, M.; Sugiyama, J.; Okano, T. Native Celluloses on the Basis of Two Crystalline

546

Phase (Iα/Iβ) System. J. Appl. Polym. Sci. 1993, 49, 1491–1496.

547

(48) Kimura, S.; Itoh, T. Cellulose Network of Hemocoel in Selected Compound Styelid

548

Ascidians. J. Electron Microsc. 1997, 46, 327‒335.

549

(49) Helbert, W.; Sugiyama, J.; Kimura, S.; Itoh, T. High-Resolution Electron Microscopy on

550

Ultrathin Sections of Cellulose Microfibrils Generated by Glomerulocytes in Polyzoa

551

vesiculiphora. Protoplasma 1998, 203, 84‒90.

552

(50) Helbert, W.; Nishiyama, Y.; Okano, T.; Sugiyama, J. Molecular Imaging of Halocynthia

553

papillosa Cellulose. J Struct. Biol. 1998, 124, 42‒50.

554

(51) Malma, E.; Bulone, V.; Wickholm, K.; Larsson, P. T.; Iversen, T. The Surface Structure

555

of Well-Ordered Native Cellulose Fibrils in Contact with Water. Carbohydr. Res. 2010, 345,

556

97–100.

557

(52) Newman, R. H.; Davidson, T. C. Molecular Conformations at the Cellulose-Water

558

Interface. Cellulose 2004, 11, 23‒32.

559

(53) Larsson, P. T.; Wickholm, K.; Iversen, T. A CP/MAS

560

Molecular Ordering in Celluloses. Carbohydr. Res. 1997, 302, 19–25.

561

(54) Viëtor, R. J.; Newman, R. H.; Ha, M. A.; Apperley, D. C.; Jarvis, M. C. Conformational

562

Features of Crystal-Surface Cellulose from Higher Plants. Plant J. 2002, 30, 721‒731.

563

(55) VanderHart, D. L.; Atalla, R. H. Studies of Microstructure in Native Celluloses Using

564

Solid-State Carbon-13 NMR. Macromolecules 1984, 17, 1465–1472.

565

(56) Lennholm, H.; Larsson, T.; Iversen, T. Determination of Cellulose Iα and Iβ in

566

Lignocellulosic Materials. Carbohydr. Res. 1994, 261, 119–131.

567

(57) Suzuki, S.; Suzuki, F.; Kanie, Y.; Tsujitani, K.; Hirai, A.; Kaji, H.; Horii, F. Structure

13

C NMR Investigation of

24 ACS Paragon Plus Environment

Page 25 of 26

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

568

and Crystallization of Sub-Elementary Fibrils of Bacterial Cellulose Isolated by Using a

569

Fluorescent Brightening Agent. Cellulose 2012, 19, 713‒727.

570 571

25 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

572

Page 26 of 26

Table of Contents

573 574

Different Conformations of Surface Cellulose Molecules in Native

575

Cellulose Microfibrils Revealed by Layer-by-Layer Peeling

576 577

Ryunosuke Funahashi, Yusuke Okita, Hiromasa Hondo, Mengchen Zhao, Tsuguyuki Saito, and

578

Akira Isogai*

579 580 581 582 583 584 585 586

26 ACS Paragon Plus Environment