morimune-moriya2018

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

Subscriber access provided by Washington University | Libraries

Reinforcement Effects from Nanodiamond in Cellulose Nanofibril Films


Seira Morimune-Moriya, Michaela Salajkova, Qi Zhou, Takashi Nishino, and Lars A. Berglund
Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.8b00010 • Publication Date (Web): 05 Apr 2018
Downloaded from http://pubs.acs.org on April 7, 2018

Just Accepted

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 34 Biomacromolecules

1
2
3
4
5
6
7
8
Reinforcement Effects from Nanodiamond in
9
10
11
12
Cellulose Nanofibril Films
13
14
15
16 Seira Morimune-Moriya1, Michaela Salajkova2,3, Qi Zhou3,4, Takashi Nishino5, Lars A.
17
18
19
Berglund2,3*
20
21
22 1. Department of Applied Chemistry, College of Engineering, Chubu University, Matsumoto,
23
24 Kasugai 487-8501 (Japan) 2. Department of Fibre and Polymer Technology, Royal Institute of
25
26
Technology, SE-100 44 Stockholm (Sweden) 3. Wallenberg Wood Science Center, Royal
27
28
29 Institute of Technology, SE-100 44 Stockholm (Sweden) 4. School of Biotechnology, Royal
30
31 Institute of Technology, AlbaNova University Centre, SE-106 91 Stockholm (Sweden) 5.
32
33 Department of chemical science and engineering, Graduate School of Engineering, Kobe
34
35
36 University, Rokko, Nada, Kobe 657-8501 (Japan)
37
38
39
40
41
42 KEYWORDS. Biocomposite, Hardness, Nanocomposite, Decoration, Strength, Modulus,
43
44
45
Nanopaper
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 2 of 34

1
2
3 ABSTRACT.
4
5
6
7 Although research on nanopaper structures from cellulose nanofibrils (CNF) is well established,
8
9 the mechanical behavior is not well understood, especially not when CNF is combined with hard
10
11 nanoparticles. Cationic CNF (Q-CNF) was prepared and successfully decorated by anionic
12
13
14
nanodiamond (ND) nanoparticles in hydrocolloidal form. The Q-CNF/ND nanocomposites were
15
16 filtered from a hydrocolloid and dried. Unlike many other carbon nanocomposites, the Q-
17
18 CNF/ND nanocomposites were optically transparent. Reinforcement effects from the
19
20
nanodiamond were remarkable, such as Young’s modulus (9.8 GPa  16.6 GPa) and tensile
21
22
23 strength (209.5 MPa  277.5 MPa) at a content of only 1.9% v/v of ND, and the reinforcement
24
25 mechanisms are discussed. Strong effects on CNF network deformation mechanisms were
26
27 revealed by loading-unloading experiments. Scratch hardness also increased strongly with
28
29
30 increased addition of ND.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 3 of 34 Biomacromolecules

1
2
3 1. Introduction
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 Figure 1 Transmission electron microscopy image of nanodiamond particles.
33
34
35 Diamond has a Young’s modulus as high as 1 TPa and a hardnessas high as 10 on the Mohs
36
37
38 scale1. Nanodiamond (ND), can be readily prepared by a detonation method and can now be used
39
40 in nanocomposites. ND nanoparticles have a diameter below 10 nm (see Figure 1 and Supporting
41
42 Info Figure 1). These ND nanoparticles can be described as core-shell particles. The diamond
43
44
45
particle forms the core, whereas the shell has oxygen containing groups2 attached to graphite.
46
47 The properties of these ND are characteristic of diamond, with the added feature of nanoscale
48
49 particle form. The characteristics of the shell makes water dispersion of ND facile, and ND can
50
51 also be modified and functionalized for specific purposes3,4. Nanodiamond is a candidate
52
53
54 nanoparticle for nanocomposites,5–8 where hardness is a desirable property.
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 4 of 34

1
2
3 In a previous study at Kobe University9, nanodiamond nanocomposites were prepared in
4
5
6 water-assisted processing. Polyvinyl alcohol (PVA) was dissolved in water and used as a matrix.
7
8 The properties of ND provided strong reinforcement of PVA in the nanocomposites, and an
9
10 important reason for this was the good dispersion of nanodiamond in PVA. For example, from
11
12
13
tensile tests, the Young’s modulus of PVA was doubled at a nanodiamond content of only 1
14
15 weight percent. One challenge in the analysis of reinforcement effects on PVA, is that the
16
17 nanoparticle will influence crystalline morphology, time-dependence of deformation behavior
18
19
and physical ageing. As a consequence, part of the measured reinforcement effect may be due to
20
21
22 ND-induced changes in PVA structure and properties10.
23
24
25 Here, a nanocomposite material is prepared from cellulose nanofibrils (CNF) reinforced by
26
27 nanodiamond particles. Cellulose is a widely available nanofibrious polymer from renewable
28
29
30 resources. Cellulose is a large research field, and it is widely used industrially. There is also
31
32 substantial growth potential due to the increasing interest in eco-friendly materials11–13. CNF is
33
34 cellulose in nanofibrillar form, and the fibrils are composed of high molar mass extended and
35
36
37
ordered cellulose molecules. From X-ray diffraction studies of plant fibers under tension, the
38
39 crystalline region of cellulose I is estimated to have an elastic modulus of around 135 GPa14,15
40
41 and it is interesting to compare this with the modulus for aluminum (70 GPa) or glass fiber (76
42
43
GPa), since the cellulose crystal density is much lower (≈1600 kg/m3). The strength of CNF
44
45
46 disintegrated from wood pulp was estimated to be a few GPa16, although fibril strength in plants
47
48 is expected to be higher (higher molar mass and fewer defects). Recently, the strength was
49
50 suggested to be comparable to multiwalled carbon nanotubes17, although this will depend on the
51
52
53 specific type of nanotube. In addition to high mechanical properties, CNF shows low thermal
54
55 expansion in the axial direction and low density. Due to the properties and nanoscale dimensions,
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 5 of 34 Biomacromolecules

1
2
3 CNF has been widely used for nanocomposites18–29. In many cases, CNF was the reinforcement
4
5
6 phase in polymer composites. For example, PVA20,21,24, epoxy resins21, polylactic acid22 and
7
8 polyurethanes23 were investigated in their roles as matrix phase.
9
10
11 Henriksson and colleagues.30 studied cellulose films (“nanopaper”) based on wood CNF. The
12
13
14
nanopaper was not solvent cast, but rather filtered from a hydrocolloidal suspension. Despite
15
16 high void content (28%), the modulus was high (≈13 GPa), the strength was high (≈210 MPa)
17
18 and work to fracture (15 MJ/m3). High optical transparency31, oxygen barrier properties32 and
19
20
favorable thermal expansion coefficient33,34 are other interesting characteristics of nanocellulose
21
22
23 films. CNF is organized as a swirled, random-in-the-plane nanofiber network, where fibrils are
24
25 strongly bonded due to small diameter (4-10nm), smooth fibril surfaces and the drying procedure.
26
27 As water evaporates during drying of cellulose nanopaper, nanometer-sized fibrils are brought
28
29
30 together facilitating molecular scale interaction, and the result is strong secondary fibril-fibril
31
32 bonding. Benitez and Walther thoroughly reviewed the mechanical properties of cellulose
33
34 nanopaper56. Effects from relative humidity are discussed as well as effects from fibril structure,
35
36
37
counterions, porosity etc.
38
39
40 An interesting possibility for increased property range and even new functionalities is to
41
42 combine cellulose nanopaper with inorganic nanoparticles. Liu et al.35 combined CNF with
43
44
montmorrillonite nanoclay for reinforcement purposes. The fire retardancy of the material was
45
46
47 dramatically improved with delayed thermal degradation of cellulose. Furthermore,
48
49 montmorrilonite strongly enhanced the gas barrier properties, also under humid conditions. Anti-
50
51 bacterial properties were achieved by silver nanocluster addition36 and zinc oxide nanoparticles37.
52
53
54 CNF/graphene nanocomposites38 showed greatly improved mechanical properties due to the high
55
56 aspect ratio of graphene platelets.
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 6 of 34

1
2
3 In the present study, another class of inorganic cellulose hybrids was investigated.
4
5
6 Nanodiamond nanocomposites were prepared by combination of cellulose nanofibrils with
7
8 nanodiamond particles, so that the hard ND particles were decorating the CNF fibrils in the
9
10 nanopaper structure. A key was to use quaternized cellulose nanofibrils (Q-CNF), so that the
11
12
13
electropositive charge on the CNF could attract the electronegative ND particles. (Supporting
14
15 Info, Table 1). First, the cationic Q-CNF was just mixed with anionic ND in hydrocolloidal state,
16
17 so that ND could attach along the CNF fibrils. The Q-CNF/ND filtered and the hybrid
18
19
nanocomposites were obtained in a similar manner as has been previously reported Sehaqui et
20
21
22 al.39 Nanostructural details of the materials were investigated, as well as the reinforcement
23
24 effects from the nanodiamond particles. The property improvements were quite remarkable at
25
26 low particle content, and possible mechanisms are discussed.
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 7 of 34 Biomacromolecules

1
2
3
4
2. Experimental
5
6
7 2.1. Materials
8
9
10
11 Q-CNF. Cationic CNF was prepared from chemical softwood pulp using glycidyl
12
13 trimethylammonium chloride (GTAC, Sigma-Aldrich, USA), and the procedure provided in Pei
14
15 et al.40 The sulphite pulp was never dried (Nordic Paper, Sweden) and subjected to beating in a
16
17
18 PFI-mill (HAM-JERN, Norway), and mixed with a sodium hydroxide (Sigma-Aldrich, USA)
19
20 solution in water. The quaternization reaction conditions were 65 °C for 8 h with stirring after
21
22 adding GTAC to the suspension. The mixture was then neutralized with hydrochloric acid (37%,
23
24
25
Sigma-Aldrich, USA) and washed with deionized water. The aqueous suspension of the
26
27 chemically treated pulp (0.5% w/v) was stirred for 24 h at 600 rpm, and passed through a
28
29 microfluidizer (M-110EH, Microfluidics Ind., USA) equipped with 200 and 100 µm chamber at
30
31 a pressure of 1600 bar at room temperature (21 °C). The Q-CNF hydrocolloid showed a solid
32
33
34 content of 0.3–0.4%.
35
36
37 The ammonium chloride group content on the treated fibrils were determined by
38
39 conductometric titration41. 100 mg (dry weight) of Q-CNF suspended in Milli-Q water (0.1%
40
41
42 w/w) was titrated with 0.005 M silver nitrate (AgNO3, Sigma-Aldrich, USA) solution by adding
43
44 200 µL. The conductivity was recorded with a conductivity meter (Mettler Toledo, USA) in 60
45
46 second intervals. The amount of trimethylammonium groups in Q-CNF can be calculated based
47
48
on the volume of AgNO3 used. The charge density of Q-CNF was 0.56 mmol/g.
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 8 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Scheme 1. Method for processing of Q-CNF/ND materials.
23
24
25 Q-CNF/ND nanocomposites. The preparation of the Q-CNF/ND nanocomposites is described
26
27
in Scheme 1. The Q-CNF suspension was added to the ND aqueous suspension (Bando Chem.
28
29
30 Ind.) to reach a Q-CNF concentration of 0.2%w/w. The ND content in the suspension was 0–5%
31
32 w/w and the rest was Q-CNF. The Q-CNF/ND suspension was stirred for 48 h and filtered on a
33
34 glass filter funnel (7.2 cm in diameter) using filter membrane (0.65 µm DVPP, Millipore, USA).
35
36
37 The Q-CNF/ND suspension was stable. After the filtration, the wet cake of the nanocomposite,
38
39 which contained 78–81% of water, was placed between the metal mesh sheets on filter paper and
40
41 then dried at 93 °C for 15 min under vacuum by using Rapid Köthen (PTI, Austria). This
42
43
44
resulted in the Q-CNF/ND nanocomposites with thickness in the range of 50–60 µm.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 9 of 34 Biomacromolecules

1
2
3
4
2.2. Characterization
5
6
7 Atomic Force Microscopy (AFM). Nanoscope IIIa AFM (Picoforce SPM, Veeco, USA) was
8
9 used to observe the Q-CNF decorated with ND. All measurements were performed in the tapping
10
11
12 mode with a scan rate of 2 Hz/512 dots using standard noncontact silicon cantilevers (RTESP,
13
14 Veeco, USA). The samples were prepared on a mica substrate by depositing the diluted Q-
15
16 CNF/ND nanocomposite suspension.
17
18
19
20
Transmission Electron Microscopy (TEM). A highly diluted Q-CNF/ND nanocomposite
21
22 suspension was deposited on a copper grid (ultra-thin carbon film/holey carbon, Ted Pella,
23
24 USA), and then stained by uranyl acetate negative stain. Excess suspension liquid was removed
25
26
using filter paper and the remaining sample was carefully dried in order to minimize the risk for
27
28
29 drying artifacts. The sample was observed at 80 kV using transmission electron microscopy
30
31 (Hitachi HT-7700, Hitachi, Japan).
32
33
34 Field-Emission Scanning Electron Microscopy (FE-SEM). The cross sections of the
35
36
37 nanocomposites were observed by FE-SEM using Hitachi S-4800 (Hitachi, Japan) equipped with
38
39 a cold field emission electron source. The samples were coated with graphite and platinum-
40
41 palladium using Cressington 208 HR sputter coaters (Cressington Scientific Instruments Ltd.,
42
43
44
UK). Secondary electron detector was used for capturing images at 3 kV/5 µA.
45
46
47 Ultraviolet-visible spectroscopy (UV-VIS). UV-VIS spectra of the nanocomposites were
48
49 observed at room temperature using UV-Visible spectrophotometer (SHIMADZU UV-2550,
50
51
Shimadzu, Japan) at a wavelength scan rate of 400 nm/min.
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 10 of 34

1
2
3 Thermogravimetric analysis (TGA). The composition of the nanocomposites was obtained by
4
5
6 TGA using Mellter-Toledo thermogravimetric analyzer (TGA/SDTA851, Switzerland). The
7
8 heating rate of 10 °C/min was used under nitrogen flow. ND content of the nanocomposites was
9
10 calculated by using the weight of residual at 550 °C, where the weight loss of CNF became
11
12
13
constant, considering the weight loss of ND.
14
15
16 Tensile test. The tensile properties were tested by Instron 5944 mechanical testing system
17
18 (Instron, USA) equipped with 500 N load cell. The specimens were conditioned at 50%RH and
19
20
23 °C. The gauge length was set at 20 mm and the cross-head speed was 2 mm/min. A minimum
21
22
23 of 6 specimens were tested for each sample. The loading-unloading test was carried out with the
24
25 same instrument. The loading-unloading cycle was repeated with the step of 1% strain.
26
27
28 Scratch test. The scratch resistance of the nanocomposites was measured by Nano scratch tester
29
30
31 (CSM Instrument, Switzerland). The test was performed with a sphero-conical diamond tip
32
33 (diameter 2 µm) and the maximum load of 10 mN was applied. The scratch map over 3 points on
34
35 each sample were made and the width of the scratch marks were measured. The scratch hardness
36
37
38
was calculated using the following equation (1)42;
39
40
ସி
41 ‫ݍ=ܪ‬ (1)
గ௪ మ
42
43
44
45 where the F is normal load (N), w is residual width of the scratch mark and q is a function of the
46
47 viscoelasticity of the material. q = 2 corresponds to the rigid plastic materials and 1 < q < 2 is for
48
49 visco-elastic materials.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 11 of 34 Biomacromolecules

1
2
3
4
3. Results and Discussion
5
6
7 3.1. Characterization of the Q-CNF decorated with ND
8
9
10
11
12
13 A BC
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 C D
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Figure 2. (A) AFM height image, (B) AFM phase image and (C, D) TEM image of decorated Q-
54
55 CNF.
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 12 of 34

1
2
3 The Q-CNF fibrils were decorated with ND as described in the experimental section. AFM
4
5
6 height and phase images of the decorated Q-CNF can be observed in Figures 2A and B. The
7
8 contrast in the phase image is high and this is due to the large difference in modulus between Q-
9
10 CNF and ND. In the phase image in Figure 2B, ND is present as distinct dots, whereas Q-CNF
11
12
13
has similar color as the background. The nanodiamond dots are apparent along the Q-CNF, and
14
15 this strengthens the hypothesis that ND is indeed adsorbed to Q-CNF. The Q-CNF and
16
17 nanodiamond distribution was also studied by transmission electron microscopy. Figure 2C
18
19
presents decorated Q-CNF in diluted Q-CNF/ND nanocomposite suspension. Although there is
20
21
22 some CNF agglomeration, the information in the image is helpful. In this figure, ND appears on
23
24 the surface of Q-CNF, in support of successful ND-decoration of Q-CNF. In contrast, the neat Q-
25
26 CNF showed a smooth surface with a width below 10 nm (Supporting figure 2) as previously
27
28
29 reported40. According to this previous study, Q-CNF dimensions are below 5 nm in diameter and
30
31 around 1 µm in length. The fibrils are flexible, so that a swirled, intermingled network structure
32
33 is obtained in dried films.
34
35
36
37
From TEM images, the distribution of ND was studied (Figure 2D and Supporting figure 3).
38
39 Q-CNF was primarily combined with agglomerates of ND with a size up to 50 nm. ND adsorbed
40
41 to Q-CNF because of the opposite charge on the fibrils and the ND nanoparticles. Images support
42
43
good small-scale dispersion of nanodiamond in the Q-CNF network.
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 13 of 34 Biomacromolecules

1
2
3
4
3.2. Structure of CNF-nanodiamond materials
5
6
7
8
A
9
10
11
12
13
14
15
16
17
18
19
B
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37 C D
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 14 of 34

1
2
3 Figure 3. (A) Photographic images for transparency impression and (B) UV-VIS spectra of Q-
4
5
6 CNF film and Q-CNF/ND nanocomposites, (C) FE-SEM images comparing cross sections of Q-
7
8 CNF reference film and (D) Q-CNF/ND nanocomposite (5% w/w).
9
10
11 Table 1. Composition and estimated porosity of Q-CNF film and Q-CNF/ND nanocomposites.
12
13
14 ND Porosity
15 Sample
(%v/v) (%)
16
17 Q-CNF 0 6.4
18 Q-CNF/ND
19 0.1 4.4
0.5%w/w
20
Q-CNF/ND
21 0.4 4.3
22 1%w/w
23 Q-CNF/ND
1.1 2.2
24 2.5%w/w
25 Q-CNF/ND
26 1.9 1.2
5%w/w
27
28
29
30
31 The optical transparency of nanocomposites based on carbon nanotubes or graphene tends to
32
33 be poor even at low nanoparticle content. The transparency of polyvinyl alcohol (PVA) was
34
35 reduced by about 35% when only 0.1 weight percent of carbon nanotubes (SWNT) were added.
36
37 43
38 Nanocomposites based on graphene and cellulose nanofibrils with a thickness below 10 µm
39
40 also lost transparency at 1.25% w/w graphene content38. In contrast, the present Q-CNF/ND
41
42 nanocomposites maintain reasonably high optical transparency, even at a concentration of 5%
43
44
w/w of ND, see Figures 3A and B. This indicates that the present composition and preparation
45
46
47 strategy provide low porosity in the nanocomposites and fairly homogeneous dispersion of ND9.
48
49
50 The content of nanodiamond was estimated from TGA and is presented in Table 1. The
51
52 measured amount of nanodiamond for low congtent nanocomposites was almost the same as the
53
54
55 starting content, while the nanocomposites with high nanodiamond concentration showed much
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 15 of 34 Biomacromolecules

1
2
3 lower content compared with the initial value (Supporting Info, Table 2). This suggests that some
4
5
6 excess free ND in the hydrocolloid was lost through the membrane, whereas ND particles
7
8 adsorbed to the Q-CNF were preserved in the nanocomposites.
9
10
11 Figure 3C shows FE-SEM images of the cross section of the Q-CNF. The structure is layered,
12
13
14
which is typical for cellulose nanopaper30. For nanocomposites, a dense structure with low
15
16 porosity was observed, see Figure 3D. Due to the filtration and drying procedure, the structural
17
18 organization of the present nanocomposite is essentially a low-porosity random-in-plane
19
20
nanofibril network with swirled, intermingled fibrils <5nm in diameter by 1 µm in length.40 It
21
22
23 was suggested that ND is located in interfibril regions, due to the Q-CNF/ND decoration
24
25 approach. This is an unusual structure, since the Q-CNF “matrix” is fibrillar in nature, and the
26
27 ND is located in a region critical to interfibril stress-transfer. Figure 3D also indicates that the
28
29
30 material is organized in sheets, a form of meso-scale material architecture.
31
32
33 In nanocomposites with very high content of ND (> 10% w/w), large agglomerates were
34
35 observed in the nanocomposite structure. Agglomerates are likely to form from clustering of
36
37
38
excess individual ND, in particular free ND which were not lost during filtration. The estimated
39
40 porosity of high ND content nanocomposites increased with increasing ND concentration
41
42 (Supporting Info Figure 4 and Supporting Info Table 2).
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 16 of 34

1
2
3
4
3.3. Mechanical properties of nanocomposites in uniaxial tension
5
6
7
8 A B
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 C
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
Figure 4. (A) Stress versus strain curves of Q-CNF film and Q-CNF/ND nanocomposites in
46
47 uniaxial tensile loading; (B) Stress versus strain curves of Q-CNF/ND nanocomposites (1% w/w)
48
49 from loading-unloading experiment; (C) Young’s modulus as a function of number of loading
50
51
steps for Q-CNF film and Q-CNF/ND nanocomposite (1% w/w) specimen tested by loading-
52
53
54 unloading. The modulus determined during loading and initial unloading are represented as
55
56 square and triangular dots, respectively.
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 17 of 34 Biomacromolecules

1
2
3 Table 2. Properties for Q-CNF nanopaper and Q-CNF/ND materials.
4
5
6
Young’s Tensile Elongation at Work of
7 Yield strength
Sample modulus strength break fracture
8 (MPa) 3
9 (GPa) (MPa) (%) (MJ/m )
10
11 Q-CNF 9.8 ± 0.9 210 ± 27 103 8.0 ± 0.7 8.2
12
13 Q-CNF/ND
10.0 ± 0.4 211 ± 17 110 8.0 ± 0.5 8.4
14 0.5%w/w
15 Q-CNF/ND
16 13.6 ± 0.8 250 ± 10 142 6.6 ± 1.3 8.1
1%w/w
17
18 Q-CNF/ND
15.6 ± 0.8 265 ± 27 147 7.3 ± 0.8 9.7
19 2.5%w/w
20 Q-CNF/ND
21
16.6 ± 1.5 278 ± 29 147 6.8 ± 0.7 9.2
5%w/w
22
23
24
25
26 Figure 4A and Table 2 present results from uniaxial tensile tests for the Q-CNF film and the
27
28 Q-CNF/ND nanocomposites. The mechanical properties of the present Q-CNF nanopaper are
29
30 lower than for films prepared from enzymatic CNF30,40. At low addition of nanodiamond, 0.5
31
32
33 weight percent the reinforcement effect was very small. Then at 1 weight percent of
34
35 nanodiamond addition, the reinforcement effect was dramatic. For instance, modulus increased
36
37 from 9.8 to 13.6 GPa and yield strength from 103 to 142 MPa. At 5% ND, the modulus was 16.6
38
39
40
GPa, and ultimate strength had increased from 210 MPa for neat CNF to 278 MPa for the
41
42 nanocomposite. The strain to failure of the Q-CNF film was preserved with ND addition,
43
44 resulting in increased work to fracture, corresponding to the area under the stress-strain curves.
45
46
47
Modulus data are examined as a function of ND content. At 0.5% w/w, the ND is not effective
48
49
50 in increasing modulus very much. At 1% w/w (0.4% v/v), the reinforcement effect is very strong,
51
52 but then the relative effect decreases with higher ND content, due to agglomeration effects. The
53
54 mechanism for strong reinforcement effects is unclear. Theoretical predictions based on the
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 18 of 34

1
2
3 Halpin-Tsai model for spherical particles in a polymer matrix44 do not predict strong
4
5
6 reinforcement effects at low particle content. The presence of ND at the interfibril interface is
7
8 significant. In interfibril stress transfer, shear deformation is important for modulus. One may
9
10 speculate that in the present system, viscoelastic deformation effects are suppressed due to
11
12
13
improved shear stress transfer as interfibril interactions are influenced by the presence of ND.
14
15 Stress relaxation effects normally included in quasi-static tensile tests may be reduced and the
16
17 material becomes strongly reinforced by the ND. Another factor, possibly related, is that the
18
19
composites have lower porosity than Q-CNF as shown in Table 145.
20
21
22
23 From Figure 4A, it is also apparent that yield strength is substantially increased. Yielding is
24
25 related to irreversible interfibril deformation,30 so the yield strength effect is in support of the
26
27 proposed reinforcement mechanism. Ultimate strength increases as a consequence of increased
28
29
30 yield strength and conserved, or even increased strain hardening coefficient in the plastic
31
32 deformation region. Strain-to-failure is similar for all materials, suggesting strain-controlled
33
34 failure.
35
36
37
38
Strongly increased mechanical properties were also reported for CNF/graphene
39
40 nanocomposites38. There were substantial improvements in the Young’s modulus, toughness and
41
42 tensile strengths. It was stated that the interface was a key reason, and that the amphiphilic nature
43
44
of CNF and the hydrophobic nature of graphene was stabilized by π-interaction. The present
45
46
47 explanation is more centered around effects of ND on CNF network deformation. The ND
48
49 decoration of the Q-CNF, may constrain network deformation, in particular for nanocomposites
50
51 where the nanodiamond content is higher than 1 weight percent. In the case of nanocomposites
52
53
54 with exceptionally high ND content (7.5, 10, 25, 50% w/w) the reinforcement efficiency was
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 19 of 34 Biomacromolecules

1
2
3 very low, most likely since ND formed large agglomerates (Supporting Info, Table 3). An
4
5
6 interesting observation is that scratch hardness increased with increased ND content (SI, Table 3).
7
8
9 Another experiment was conducted by loading the material, followed by unloading, and this
10
11 was repeated with increasing levels of strain. In Figure 4B, stress versus strain diagrams of Q-
12
13
14
CNF/ND nanocomposites (1% w/w) are presented. The modulus determined during loading and
15
16 initial unloading for Q-CNF nanopaper and Q-CNF/ND materials (1% w/w) is presented in
17
18 Figure 4C. The modulus of the Q-CNF film increased with increased strain during loading steps,
19
20
as previously observed for CNF nanopaper from enzymatic CNF30. The reason is increased
21
22
23 orientation of the CNF fibrils in the network. For the nanocomposite, the yield strength is
24
25 increased with increasing plastic strain, possibly due to increased CNF orientation. Interestingly,
26
27 the nanocomposite modulus showed stronger increase than for the Q-CNF nanopaper.
28
29
30 Henriksson et al.30 proposed that increased Young’s modulus and yield strength are due to
31
32 network and CNF fibril orientation effects in the deformation direction46. One possibility is that
33
34 the presence of ND nanoparticles is increasing the degree of Q-CNF orientation at a given strain,
35
36
37
although the mechanism for such an effect is unclear.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 20 of 34

1
2
3
4
3.4. Scratch resistance in Q-CNF/ND materials
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Figure 5. Hardness obtained by model predictions (Equation (2) and (3)) as well as experimental
28
29 scratch hardness data (Experimental) for Q-CNF/ND nanocomposites as a function of ND
30
31 content.
32
33
34 Table 3. Scratch hardness of Q-CNF film and Q-CNF/ND nanocomposites.
35
36
37
Scratch hardness
38 Sample
(MPa)
39
40
Q-CNF 272 ± 88
41
42
43 Q-CNF/ND 0.5%w/w 302 ± 73
44
45 Q-CNF/ND 1%w/w 344 ± 92
46
47
48 Q-CNF/ND 2.5%w/w 352 ± 69
49
50 Q-CNF/ND 5%w/w 466 ± 64
51
52
53
54 Experiments were performed as described in experimental section. The scratch hardness is
55
56 based on the width of the scratch mark. If the scratch mark is more narrow, then the scratch
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 21 of 34 Biomacromolecules

1
2
3 hardness and the scratch resistance is increased47. Increased scratch hardness is often expected
4
5
6 when inorganic fillers are added to polymers48,49. Although the hardness of the filler is a key
7
8 parameter, the filler can also act as a nucleation agent so that the intrinsic polymer properties are
9
10 improved by increased crystallinity, resulting in improved scratch resistance50. Neizel et al.51
11
12
13
presented scratch resistance data for epoxy/nanodiamond materials of high ND content. The ND
14
15 particles formed a connected network structure There was a decrease in penetration depth by 1.5
16
17 µm at 25 weight percent nanodiamond content. Zhang et al.52 showed 1% w/w acyl chloride
18
19
functionalized ND enhanced the hardness of polyimide by ~30%. In this system, it was
20
21
22 suggested that due to the modification of ND, strong chemical bonding was produced between
23
24 polyimide and ND contributing to the significant increase in hardness.
25
26
27 The hardness can be roughly estimated from the Young’s modulus and the yield strength. The
28
29
30 following equations (2) and (3) were used53,54:
31
32
ଶ ா ୲ୟ୬ ఉ
‫ߪ × = ܪ‬௬ ቆ1 + ݈݊ ൬ ൰ቇ
33
(2)
34 ଷ ଷ(ଵିఔ మ )ఙ೤
35
36
37
ଶ ா ୡ୭ୱ ఏ
‫ߪ × = ܪ‬௬ ቆ1 + ݈݊ ൬ ൰ቇ
38
39 ଷ ଷఙ೤
(3)
40
41
42
43 where H is hardness, σy is yield strength, E is Young’s modulus and ν is Poisson’s ratio. β and θ
44
45 correspond to the indenter angles.
46
47
48 Note that this hardness is not strictly equal to scratch hardness, so the comparison is rough, for
49
50
51
the purpose of identifying underlying physical properties of the material. Figure 5 shows the
52
53 predicted hardness and experimental data for scratch hardness for the Q-CNF/ND materials with
54
55 increasing nanodiamond fraction. Predictions and experimental data show agreement up to 1
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 22 of 34

1
2
3 weight percent nanodiamond content. The hardness of the material is increasing since the
4
5
6 nanostructured composite has well-dispersed nanodiamond particles. The width of the scratch
7
8 mark is decreased in materials with nanodiamond. At 1 weight percent of nanodiamond, the
9
10 width was decreased by 1 µm and as a consequence, the scratch hardness increased by 26%, see
11
12
13
Table 3. The larger difference between experimental hardness and predicted data for the
14
15 material with 5 weight percent nanodiamond is due to the less favorable dispersion of ND at
16
17 higher reinforcement contents55. Interestingly, the scratch hardness of the Q-CNF/ND
18
19
nanocomposites was increased by further addition of ND (> 10% w/w) while the Young’s
20
21
22 modulus and the yield strength decreased (Supporting Table 3).
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 23 of 34 Biomacromolecules

1
2
3
4
4. Conclusions
5
6
7 A nanocomposite based on Q-CNF nanofibrils decorated with ND was prepared by a simple
8
9 paper-making process using an aqueous suspension. The Q-CNF was successfully decorated by
10
11
12 ND. In the nanocomposite structure, ND nanoparticles appeared to reduce the slight porosity of
13
14 the Q-CNF network. Despite the incorporation of nanocarbon material, the Q-CNF/ND
15
16 nanomaterials were optically transparent since the nanodiamond particles were well dispersed in
17
18
19
a low-porosity material. ND nanoparticles are embedded at the interfibril interface in a fibril
20
21 “matrix” of very long (≈1µm) random-in-the-plane swirled Q-CNF nanofibrils, below 5 nm in
22
23 diameter, forming an unusual type of nanocomposite material.
24
25
26
The reinforcement effect from addition of nanodiamond particles by ionic interaction to
27
28
29 nanofibrils and cellulose nanopaper was very strong, and ductility was preserved. For instance, at
30
31 1% of ND, modulus increased from 9.8 to 13.6 GPa and yield strength from 103 to 142 MPa. At
32
33 5% ND, the ultimate strength increased from 210 MPa for neat CNF to 278 MPa for the
34
35
36 nanocomposite.. The modulus increase cannot be predicted by conventional micromechanics
37
38 models. It may be speculated that the ND nanoparticles located at fibril-fibril interfaces, are
39
40 constraining the viscoelastic nature of CNF network deformation, and thus causing a stronger
41
42
43
reinforcement effect than expected. Loading-unloading experiments in the strain-hardening
44
45 plastic deformation region, revealed stronger orientation effects in the Q-CNF/ND
46
47 nanocomposites, compared with the neat Q-CNF cellulose nanopaper. This supports that there
48
49
are specific effects of interfacially located ND on the deformation mechanisms in the Q-CNF
50
51
52 network “matrix”. The scratch hardness was also strongly increased at ND contents up to 1
53
54 percent by weight, and comparable with rough theoretical estimates of hardness. The hardness
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 24 of 34

1
2
3 was also increased at very high ND content (5, 7.5, 10, 25, 50 wt%). The increase continued
4
5
6 although ND agglomeration effects were significant. Strength and modulus data did not increase
7
8 beyond 7.5wt% ND content.
9
10
11 In summary, by attaching ND ionically to Q-CNF, the unique properties of ND were
12
13
14
synergetically exploited for strongly improved properties of Q-CNF/ND hybrid nanomaterials.
15
16 The present nanocomposite is of renewable resource base, it is strong, exceptionally stiff and
17
18 optically transparent, at very low content of ND (0.1-1.9% v/v). Most “homeopathic”
19
20
reinforcement effects reported for nanocomposites in literature are due to significant changes in
21
22
23 the polymer matrix structure compared with the neat polymer reference, such as increased
24
25 crystallinity or orientation. The present “matrix” is unchanged in structure. It is therefore an
26
27 interesting system for further studies on reinforcement effects in this class of nanoparticle
28
29
30 composites.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 25 of 34 Biomacromolecules

1
2
3 Supporting Information.
4
5
6
7 Figures showing size distribution of ND in aqueous suspension, STEM image and AFM images
8
9 of Q-CNF from the diluted suspension, TEM image of distribution of ND in Q-CNF/ND
10
11 nanocomposite suspension, FE-SEM image of the cross section of Q-CNF/ND nanocomposite
12
13
14
(25% w/w). Table presenting zeta potential of ND, composition (TGA) and the porosity of Q-
15
16 CNF film and Q-CNF/ND nanocomposites and table summarizing mechanical properties and
17
18 scratch hardness of Q-CNF film and Q-CNF/ND nanocomposites. This material is available free
19
20
of charge via the Internet at http://pubs.acs.org.
21
22
23
24
25
26 AUTHOR INFORMATION
27
28 Corresponding Author
29
30 *Tel: 46-8-7908118, Fax: 46-8-7908108, E-mail: blund@kth.se
31
32
33
34 Notes.
35
36 The authors declare no competing financial interest.
37
38
39 ACKNOWLEDGMENT
40
41 This work was supported by Grant-in-Aid for Japan Society for the Promotion of Science (JSPS)
42
43 Fellows. We are grateful to staff at WWSC and in the Biocomposites group at KTH for technical
44
45
46
help and the thoughtful discussions.
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 26 of 34

1
2
3 References
4
5
6
(1) Kidalov, S. V.; Shakhov, F. M.; Vul', A. Y. Thermal conductivity of nanocomposites based
7
8
9 on diamonds and nanodiamonds. Diamond Relat. Mater. 2007, 16, 2063–2066.
10
11
12 (2) Osawa, E. Recent progress and perspectives in single-digit nanodiamond. Diamond Relat.
13
14 Mater. 2007, 16, 2018–2022.
15
16
17 (3) Meinhardt, T.; Lang , D.; Dill, H.; Krueger, A. Pushing the Functionality of Diamond
18
19
20 Nanoparticles to New Horizons: Orthogonally Functionalized Nanodiamond Using Click
21
22 Chemistry. Adv. Funct. Mater. 2011, 21, 494–500.
23
24
25 (4) Wuest, K. N. R.; Trouillet, V. A.; Goldmann, S.; Stenzel, M. H.; Barner-Kowollik, C.
26
27 Polymer Functional Nanodiamonds by Light-Induced Ligation. Macromolecules 2016, 49,
28
29
30 1712−1721.
31
32
33 (5) Mochalin, V. N.; Shenderova, O.; Ho, D.; Gogotsi, Y. The properties and applications of
34
35 nanodiamonds. Nat. nanotechnol. 2012, 7, 11–23.
36
37
38 (6) Rej, E.; Gaebel, T.; Waddington, D. E. J.; Reilly, D. J. Hyperpolarized Nanodiamond
39
40
41
Surfaces. J. Am. Chem. Soc., 2017, 139, 193–199.
42
43
44 (7) Lee, D.-K.; Kim, S. V.; Limansubroto, A. N.; Yen, A.; Soundia, A.; Wang, C.-Y.; Shi, W.;
45
46 Hong, C.; Tetradis, S.; Kim, Y.; Park, N.-H.; Kang, M. K.; Ho, D. Nanodiamond–Gutta Percha
47
48 Composite Biomaterials for Root Canal Therapy. ACS Nano 2015, 9, 11490–11501.
49
50
51 (8) Zhang, F.; Song, Q.; Huang, X.; Li, F.; Wang, K.; Tang, Y.; Hou, C.; Shen H. A Novel
52
53
54 High Mechanical Property PLGA Composite Matrix Loaded with Nanodiamond–Phospholipid
55
56 Compound for Bone Tissue Engineering. ACS Appl. Mater. Interfaces 2016, 8, 1087–1097.
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 27 of 34 Biomacromolecules

1
2
3 (9) Morimune, S.; Kotera, M.; Nishino, T.; Goto, K.; Hata, K. Poly(vinyl alcohol)
4
5
6 Nanocomposites with Nanodiamond Macromolecules 2011, 44, 4415–4421.
7
8
9 (10) Roumeli, E.; Pavlidou, E.; Avgeropoulos, A.; Vourlias, G.; Bikiaris, D. N.; Chrissafis, K.
10
11 Factors Controlling the Enhanced Mechanical and Thermal Properties of Nanodiamond-
12
13 Reinforced Cross-Linked High Density Polyethylene. J. Phys. Chem. B 2014, 118, 11341–
14
15
16 11352.
17
18
19 (11) Berglund, L. A.; Peijs, T. Cellulose biocomposites-from bulk moldings to nanostructured
20
21 systems. Mrs Bull. 2010, 35, 201–207.
22
23
24 (12) Eichorn, S. J.; Dufresne, A.; Aranguren, M.;Marcovich, N. E.; Capadona, J. R.; Rowan,
25
26
27
S. J.; Weder, C.; Thielemans, W.; Roman, M.; Renneckar, S.; Gindl, W.; Veigel, S.; Keckes, J.;
28
29 Yano, H.; Abe, K.; Nogi, M.; Nakagaito, A. N.; Mangalam, A.; Simonsen, J.; Benight, A. S.;
30
31 Bismarck, A.; Berglund, L. A.; Peijs, T. Review: current international research into cellulose
32
33
nanofibres and nanocomposites. J. Mater. Sci. 2010, 45, 1–33.
34
35
36
37
(13) Klemm, D.; Kramer, F.; Moritz, S.; Lindstrm, T.; Ankerfors, M.; Gray, D.; Dorris A.
38
39 Nanocelluloses: A New Family of Nature-Based Materials. Angew. Chem. Int. Edit. 2011, 50,
40
41 5438–5466.
42
43
44 (14) Sakurada, I.; Nukushina, Y.; Ito, T. Experimental determination of the elastic modulus of
45
46
47 crystalline regions in oriented polymers. J. Polym. Sci. 1962, 57, 651–660.
48
49
50 (15) Nishino, T.; Takano, K.; Nakamae, K. Elastic Modulus of the Crystalline Regions of
51
52 Cellulose. Polymorphs. J. Polym. Sci, Part B. 1995, 33, 1647–1651.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 28 of 34

1
2
3 (16) Page, D. H.; EL-Hosseny, F. The mechanical properties of single wood pulp fibres. J.
4
5
6 Pulp Paper Sci. 1983, 9, 99–100.
7
8
9 (17) Saito, T.; Kuramae, R.; Wohlert, J.; Berglund, L. A.; Isogai, A. An Ultrastrong
10
11 Nanofibrillar Biomaterial: The Strength of Single Cellulose Nanofibrils Revealed via Sonication-
12
13 Induced Fragmentation. Biomacromolecules 2013, 14, 248−253.
14
15
16
17
(18) Samir, M.; Alloin, F.; Paillet, M.; Dufresne, A. Tangling Effect in Fibrillated Cellulose
18
19 Reinforced Nanocomposites. Macromolecules 2004, 37, 4313–4316.
20
21
22 (19) Nakagaito, A. N.; Yano, H. Appl. Phys. A: Mater. Sci. Process. 2004, 78, 547–552.
23
24
25 (20) Zimmermann, T.; Pohler, E.; Geiger, T. Adv. Eng. Mater. 2004, 6, 754–761.
26
27
28 (21) Bruce, D. M.; Hobson, R. N.; Farrent, J. W.; Hepworth, D. G. Composites, Part A 2005,
29
30
31
36, 1486–1493.
32
33
34 (22) Iwatake, A.; Nogi, M.; Yano, H. The effect of morphological changes from pulp fiber
35
36 towards nano-scale fibrillated cellulose on the mechanical properties of high-strength plant fiber
37
38 based composites. Compos. Sci. Technol. 2008, 68, 2103–2106.
39
40
41 (23) Seydibeyoglu, M. O.; Oksman, K. Novel nanocomposites based on polyurethane and
42
43
44 micro fibrillated cellulose. Compos. Sci. Technol. 2008, 68, 908–914.
45
46
47 (24) Gea, S.; Bilotti, E.; Reynolds, C. T.; Soykeabkeaw, N.; Peijs, T. Bacterial cellulose–poly
48
49 (vinyl alcohol) nanocomposites prepared by an in-situ process. Mater. Lett. 2010, 64, 901–904.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 29 of 34 Biomacromolecules

1
2
3 (25) Sehaqui, H.; Morimune, S.; Nishino, T.; Berglund, L. A. Stretchable and Strong Cellulose
4
5
6 Nanopaper Structures Based on Polymer-Coated Nanofiber Networks: An Alternative to
7
8 Nonwoven Porous Membranes from Electrospinning. Biomacromolecules 2012, 13, 3661–3667.
9
10
11 (26) Sehaqui, H.; Mushi, N. E.; Morimune, S.; Salajkova, M.; Nishino, T.; Berglund, L. A.
12
13 Cellulose nanofiber orientation in nanopaper and nanocomposites by cold drawing. ACS Appl.
14
15
16 Mater. Interfaces 2012, 4, 1043–1049.
17
18
19 (27) Hayase, G.; Kanamori, K.; Abe, K.; Yano, H.; Maeno, A.; Kaji, H.; Nakanishi, K.
20
21 Polymethylsilsesquioxane–Cellulose Nanofiber Biocomposite Aerogels with High Thermal
22
23 Insulation, Bendability, and Superhydrophobicity. ACS Appl. Mater. Interfaces 2014, 6, 9466–
24
25
26 9471.
27
28
29 (28) Yu, P.; He, H.; Luo, Y.; Jia, D.; Dufresne, A. Reinforcement of Natural Rubber: The Use
30
31 of in Situ Regenerated Cellulose from Alkaline–Urea–Aqueous System. Macromolecules 2017,
32
33
50, 7211–7221.
34
35
36
37
(29) Yao, K.; Huang, S.; Tang, H.; Xu, Y.; Buntkowsky, G.; Berglund, L. A.; Zhou, Q.
38
39 Bioinspired Interface Engineering for Moisture Resistance in Nacre-Mimetic Cellulose
40
41 Nanofibrils/Clay Nanocomposites. ACS Appl. Mater. Interfaces 2017, 9, 20169–20178.
42
43
44 (30) Henriksson, M.; Berglund, L. A.; Isaksson, P.; Lindström, T.; Nishino, T. Cellulose
45
46
47 nanopaper structures of high toughness. Biomacromolecules 2008, 9, 1579–1585.
48
49
50 (31) Nogi, M.; Iwamoto, S.; Nakagaito, A. N.; Yano, H. Optically Transparent Nanofiber
51
52 Paper. Adv. Mater. 2009, 21, 1595–1598.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 30 of 34

1
2
3 (32) Fukuzumi, H.; Saito, T.; Wata, T.; Kumamoto, Y.; Isogai, A. Transparent and high gas
4
5
6 barrier films of cellulose nanofibers prepared by TEMPO-mediated oxidation.
7
8 Biomacromolecules 2009, 10, 162–165.
9
10
11 (33) Yano, H.; Sugiyama, J.; Nakagaito, A. N.; Nogi, M.; Matsuura, T.; Hikita, M.; Handa, K.
12
13 Optically Transparent. Composites Reinforced with Networks of Bacterial Nanofibers. Adv.
14
15
16 Mater. 2005, 17, 153–155.
17
18
19 (34) Okahisa, Y.; Yoshida, A.; Miyaguchi, S.; Yano, H. Optically transparent wood- cellulose
20
21 nanocomposite as a base substrate for flexible organic light-emitting diode displays. Compos.
22
23
Sci. Technol. 2009, 69, 1958–1961.
24
25
26
27
(35) Liu, A.; Walther, A.; Ikkala, O.; Belova, L.; Berglund, L. A. Clay. Nanopaper with tough
28
29 cellulose nanofiber matrix for fire retardancy and gas barrier functions. Biomacromolecules
30
31 2011, 12, 633–641.
32
33
34 (36) Dı ´ ez, I.; Eronen, P.; Österberg, M.; Linder, M. B.; Ikkala, O.; Ras, R. H.
35
36
37 Functionalization of nanofibrillated cellulose with silver nanoclusters: fluorescence and
38
39 antibacterial activity. Macromol. Biosci. 2011, 11, 1185–1191.
40
41
42 (37) Martins, N. C. T.; Freire, C. S. R.; Neto, C. P. l.; Silvestre, A. J. D.; Causio, J.; Baldi, G.;
43
44
45 Sadocco, P.; Trindade, T. Antibacterial paper based on composite coatings of nanofibrillated
46
47 cellulose and ZnO. Colloids Surf. A 2013, 417, 111–119.
48
49
50 (38) Malho, J. M.; Laaksonen, P.; Walther, A.; Ikkala, O.; Linder, M. B. Facile Method for
51
52 Stiff, Tough, and Strong Nanocomposites by Direct Exfoliation of Multilayered Graphene into
53
54
55 Native Nanocellulose Matrix. Biomacromolecules 2012, 13, 1093−1099.
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 31 of 34 Biomacromolecules

1
2
3 (39) Sehaqui, H.; Liu, A.; Zhou, Q.; Berglund, L. A. Fast Preparation Procedure for Large, Flat
4
5
6 Cellulose and Cellulose/Inorganic Nanopaper Structures. Biomacromolecules 2010, 11, 2195–
7
8 2198.
9
10
11 (40) Pei, A.; Butchosa, N.; Berglund, L. A.; Zhou, Q. Surface quaternized cellulose nanofibrils
12
13 with high water absorbency and adsorption capacity for anionic dyes. Soft Matter 2013, 9, 2047–
14
15
16 2055.
17
18
19 (41) Hasani, M.; Cranston, E. D.; Westman, G.; Gray, D. G. Cationic surface functionalization
20
21 of cellulose nanocrystals. Soft Matter 2008, 4, 2238–2244.
22
23
24 (42) Xiang, C; Chu, J; Masuda, K. Polym. Eng. Sci. 2001, 41, 23–31.
25
26
27 (43) Li, C.; Lv, X.: Dai, J.: Cui, J.: Yan, Y. Synthesis of water-soluble single-walled carbon
28
29
30 nanotubes and its application in poly (vinyl alcohol) composites. Polym. Adv. Technol. 2013, 24,
31
32 376–382.
33
34
35 (44) Halpin, J. C.; Kardos, J. L. The Halpin-Tsai equations: A review. Polym. Eng. Sci. 1976,
36
37 16, 344–352.
38
39
40
41
(45) Henriksson, M.; Berglund L. A. Structure and Properties of Cellulose Nanocomposite
42
43 Films Containing Melamine Formaldehyde. J. Appl. Polym. Sci. 2007, 106, 2817–2824.
44
45
46 (46) Gindl, W.; Martinschitz, K. J.; Boesecke, P.; Keckes, J. Structural changes during tensile
47
48 testing of an all-cellulose composite by in situ synchrotron X-ray diffraction. Comp. Sci.
49
50
51
Technol. 2006, 66, 2639–2647.
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 32 of 34

1
2
3 (47) Surampadi, N. L.; Pesacreta, T. C.; Misra, R. D. K. The determining role of scratch
4
5
6 indenter radius on surface deformation of high density polyethylene and calcium carbonate-
7
8 reinforced composite. Mater. Sci. Eng., A 2007, 456, 218–229.
9
10
11 (48) Hadal, R.; Dasari, A.; Rohrmann, J.; Misra, R. D. K. Susceptibility to scratch surface
12
13 damage of wollastonite- and talc-containing polypropylene micrometric composites. Mater. Sci.
14
15
16 Eng., A 2004, 380, 326–339.
17
18
19 (49) Misra, R. D. K.; Hadal, R.; Duncan, S. J. Surface damage behavior during scratch
20
21 deformation of mineral reinforced polymer composites. Acta Materialia 2004, 52, 4363–4376.
22
23
24 (50) Yuan, Q.; Ramisetti, N.; Misra, R. D. K. Nanoscale near-surface deformation in polymer
25
26
27
nanocomposites. Acta Materialia 2008, 56, 2089–2100.
28
29
30 (51) Neitzel, I.; Mochalin, V.; Knoke, I.; Palmese, G. R.; Gogotsi, Y. Mechanical properties of
31
32 epoxy composites with high contents of nanodiamond. Compos. Sci. Technol. 2011, 71, 710–
33
34 716.
35
36
37 (52) Zhang, Q.; Naito, K.; Tanaka, Y.; Kagawa, Y. Grafting Polyimides from Nanodiamonds.
38
39
40 Macromolecules 2008, 41, 536–538.
41
42
43 (53) Bao, Y. W.; Wang, W.; Zhou, Y. C. Investigation of the relationship between elastic
44
45 modulus and hardness on depth-sensing indentation measurements. Acta Mater. 2004, 52, 5397–
46
47
404.
48
49
50
51
(54) Wredenberg, F.; Larsson, P. L. Scratch testing of metals and polymers: Experiments and
52
53 numerics. Wear 2009, 266, 76–83.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Page 33 of 34 Biomacromolecules

1
2
3 (55) Ayatollahi, M. R.; Alishahi, E.; Doagou-R, S.; Shadlou, S. Tribological and mechanical
4
5
6 properties of low content nanodiamond/epoxy nanocomposites. Composites, Part B 2012, 43,
7
8 3425–3430.
9
10
11 (56) Benitez, A.J.; Walther, A. Cellulose nanofibril nanopapers and bioinspired
12
13 nanocomposites: a review to understand the mechanical property space. J. Mater. Chem. A 2017,
14
15
16 5, 16003-16024.
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils
Biomacromolecules Page 34 of 34

1
2
3 For Table of Contents Use Only
4
5
6
7
8
9
10
Strong and Stiff Nanocomposites based on Cellulose
11
12
13
14 Nanofibrils Decorated with Nanodiamond
15
16
17
18
19
Seira Morimune-Moriya1, Michaela Salajkova2,3, Qi Zhou3,4, Takashi Nishino5, Lars A.
20
21 Berglund2,3*
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
cellulose nanofibrils

You might also like