Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

nanomaterials

Article
Preparation and Characterization of
Polyvinylpyrrolidone/Cellulose
Nanocrystals Composites
Marina Voronova 1 , Natalia Rubleva 1 , Nataliya Kochkina 1 , Andrei Afineevskii 2 ,
Anatoly Zakharov 1 and Oleg Surov 1, *
1 G.A. Krestov Institute of Solution Chemistry of the Russian Academy of Sciences, 1 Akademicheskaya St.,
Ivanovo 153045, Russia; miv@isc-ras.ru (M.V.); rublevanw@yandex.ru (N.R.); nek@isc-ras.ru (N.K.);
agz@isc-ras.ru (A.Z.)
2 Department of Physical and Colloid Chemistry, Ivanovo State University of Chemistry and Technology,
7 Sheremetevsky Prospect, Ivanovo 153000, Russia; afineevskiy@mail.ru
* Correspondence: ovs@isc-ras.ru; Tel.: +7-4932-351545

Received: 7 November 2018; Accepted: 4 December 2018; Published: 5 December 2018 

Abstract: Composite films and aerogels of polyvinylpyrrolidone/cellulose nanocrystals (PVP/CNC)


were prepared by solution casting and freeze-drying, respectively. Investigations into the PVP/CNC
composite films and aerogels over a wide composition range were conducted. Thermal stability,
morphology, and the resulting reinforcing effect on the PVP matrix were explored. FTIR, TGA, DSC,
X-ray diffraction, SEM, and tensile testing were used to examine the properties of the composites.
It was revealed PVP-assisted CNC self-assembly that produces uniform CNC aggregates with a
high aspect ratio (length/width). A possible model of the PVP-assisted CNC self-assembly has been
considered. Dispersibility of the composite aerogels in water and some organic solvents was studied.
It was shown that dispersing the composite aerogels in water resulted in stable colloidal suspensions.
CNC particles size in the redispersed aqueous suspensions was near similar to the CNC particles size
in never-dried CNC aqueous suspensions.

Keywords: cellulose nanocrystals; nanocomposites; self-assembly; dispersibility

1. Introduction
Cellulose nanocrystals (CNC) are crystalline nano-sized rod-like particles, that can be easily
produced from natural renewable cellulosic materials by controlled acid hydrolysis. When sulfuric
acid is used, CNC form stable aqueous suspensions due to negatively charged sulfate ester groups at
the CNC particle surface which cause electrostatic repulsion between the rod-like colloidal particles.
The diameter of the anisotropic CNC particles is from 10 to 20 nm, the length is from 100 to 300 nm,
depending on processing conditions and the raw material [1]. The main features that promote the
use of CNC are an anisotropic particle shape, a rather large surface area, a low density, a very high
modulus of elasticity and a significant reinforcing effect at a low content in a composite. As one of
the most plentiful natural polymers with superb mechanical properties, excellent biodegradability
and biocompatibility, nanocrystalline cellulose has been considered as a perfect reinforcing element
for low-cost nanocomposites [2,3]. The properties of CNC (capacity to chemical modification of the
surface hydroxyl groups, high mechanical strength, formation of a chiral nematic liquid-crystalline
phase) attract considerable attention of researchers designing new functional materials [4–9].
These days the most studies on CNC composites relate to materials with low CNC content because
of the challenge to prevent CNC agglomeration at high CNC content. Another issue is a choice of
suitable polymer matrix providing high CNC content, nano-structural control, and interface tailoring

Nanomaterials 2018, 8, 1011; doi:10.3390/nano8121011 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 1011 2 of 21

at molecular scale level. In this regard, model systems consisting of crystalline (CNC) and amorphous
(polymer matrix) components can provide a mimetic approach for research of natural phenomena [10].
In recent years, intensive efforts have been provided to design new eco-friendly polymeric materials
from renewable sources as well as hybrid structures comprising nanoscale natural biodegradable and
biocompatible polymers and inorganic materials [11,12].
Polyvinylpyrrolidone (PVP), a water-soluble, non-toxic, non-ionic amorphous polymer with
high solubility in polar solvents, has been widely used in nanoparticles synthesis [13]. Due to the
amphiphilic nature, PVP can affect morphology and nanoparticle growth by ensuring solubility in
various solvents, discriminatory surface stabilization, controlled crystal growth, playing the role of a
shape-control agent and facilitating the growth of specific crystal faces while preventing others [14–16].
PVP, with its versatile features such as a lack of toxicity, film formation, water solubility,
and adhesive power, is one of the most promising polymers for nanogels preparation [17–21]. PVP can
serve as a nanoparticle dispersant, growth modifier, surface stabilizer, hampering the agglomeration
of nanoparticles through the repulsive forces that arise from its hydrophobic carbon chains interacting
with each other in a solvent. PVP has been thought as a perspective material with a very good
film-forming ability for potential application in production of cosmetics, printing inks, detergents,
paints, coatings, shielding devices, adhesives, plastics, medicine, and pharmaceuticals [22–25].
Wu et al. [26] developed a method of polypyrrole (PPy) polymerization on CNC coated with
PVP. They achieved high conductivity for the PPy/PVP/CNC composite with an excellent specific
capacitance and improved cycling stability. The authors emphasize that physical adsorption of PVP
onto the CNC surface plays a critical role in ensuring uniform PPy coating. PVP makes the CNC
surface hydrophobic and, thus, much more appropriate for the growth of the hydrophobic PPy cover
that acts as a steric stabilizer stopping further aggregation of the nanoparticles.
Going et al. [27] reported a method for achieving dispersion of the freeze-dried CNC/PVP
composite in a water–methanol system using magnetic stirring and sonication. The effects of CNC
loading and preparation method on dispersion, solution viscosity, and mechanical properties of
electrospun PVP/CNC nanocomposites were inspected.
Huang et al. [28] produced PVP/CNC/Ag nanocomposite fibers via electrospinning and studied
their properties. The authors showed that improved strength and antimicrobial characteristics
of PVP/CNC/Ag electrospun composite fibers made the material a perspective candidate for a
biomedical application.
Chitosan/PVP/CNC composite films were prepared by Hasan et al. [29] for the development of
an efficient sustained drug release in wound dressing application. It was revealed that incorporation
of CNC in films enhanced swelling, thermal, and mechanical properties and improved resistance to
enzymatic degradation.
Recently, Gao and Jin [30] manufactured iridescent chiral nematic CNC/PVP nanocomposite
films for fast and easy distinguishing similar organic solvents, such as halogenated hydrocarbons,
skeletal isomers, and homologues. They showed that the CNC/PVP nanocomposite film could work
as a sensor for the solvents recognition due to an apparent structural color change.
Here, we hypothesize that PVP adsorption onto CNC can block lateral interactions between
the CNC and prevents their agglomeration in the lateral direction, i.e., hinders growth of the CNC
particles width upon the concentration increase or drying of the composites. On the other hand,
we suppose that freezing of CNC suspensions can align rod-like CNC particles in direction of ice
crystal growth allowing formation of the CNC aggregates with a high aspect ratio. We assume also
that these aggregates can be broken down easily in water and some organic solvents, providing good
dispersibility of the composites.
The objective of this study is to produce and characterize PVP/CNC composite films and
aerogels over a wide composition range, investigate the composites morphology and PVP-assisted
CNC self-assembly as well as explore dispersibility of the PVP/CNC aerogels in water and some
organic solvents.
Nanomaterials 2018, 8, 1011 3 of 21

Nanomaterials
2. Materials2018,
and8,Methods
x FOR PEER REVIEW 3 of 21

2.1. Materials
2.1.
Microcrystalline cellulose (MCC) (~20 micron, powder) and polyvinylpyrrolidone
polyvinylpyrrolidone (powder,
average Mw
average Mw40,000)
40,000) (Figure
(Figure 1) were
1) were purchased
purchased from Sigma-Aldrich
from Sigma-Aldrich (SaintMO,
(Saint Louis, Louis, MO,
USA). USA).
Propanol,
Propanol,and
dioxane, dioxane, and chloroform
chloroform of grade
of analytical analytical
weregrade were purchased
purchased from Sigma-Aldrich
from Sigma-Aldrich Rusused
Rus and were and
were used without further purification (Moscow, Russia). Deionized water (18 mΩ,
without further purification (Moscow, Russia). Deionized water (18 mΩ, Millipore systems) was Millipore
systems)
used was used
throughout thethroughout
experiment. the Sulfuric
experiment.
acid Sulfuric acid pure,
(chemically (chemically
GOST pure, GOST
4204-77) was4204-77) was
purchased
purchased from Chimmed (Moscow, Russia). Potassium bromide (FT-IR
from Chimmed (Moscow, Russia). Potassium bromide (FT-IR grade, >99%) was purchased fromgrade, >99%) was
purchased fromRus
Sigma-Aldrich Sigma-Aldrich Rus (Moscow, Russia).
(Moscow, Russia).

( CH2 CH )n
N O

Figure 1. Structural
Figure 1. Structural formula
formula of
of polyvinylpyrrolidone.
polyvinylpyrrolidone.

2.2. Preparation of CNC


2.2. Preparation of CNC
CNC were produced from MCC using 62 wt% H2 SO4 hydrolysis as described earlier [31].
CNC were produced from MCC using 62 wt% H2SO4 hydrolysis as described earlier [31]. The
The prepared CNC had a rod-like morphology with an average width of 15–20 nm and length of
prepared CNC had a rod-like morphology with an average width of 15–20 nm and length of 100–150
100–150 nm (aspect ratio of about 7) from transmission electron microscopy (TEM) analysis (Figure S1,
nm (aspect ratio of about 7) from transmission electron microscopy (TEM) analysis (Figure S1, Table
Table S1).
S1).
2.3. Preparation of PVP/CNC Composites
2.3. Preparation of PVP/CNC Composites
PVP/CNC composites were prepared by solution casting method similar as described earlier [32].
PVP/CNC
In brief, the PVPcomposites
solution was were prepared
prepared by by solution 0.5
dissolving casting method
g of PVP in 10similar
mL of as described
deionized earlier
water at
[32]. In brief, the PVP solution was prepared by dissolving 0.5 g of PVP in 10
room temperature for 2 h under agitation. The PVP/CNC ratio was controlled by stoichiometric mL of deionized water
at room temperature
addition for 2 h under
of a CNC suspension agitation.
to a fixed The of
quantity PVP/CNC ratio wasfollowed
the PVP solution controlledby by stoichiometric
vigorous stirring
addition of a CNC suspension to a fixed quantity of the PVP solution followed by
for 1 h. The solutions were cast into glass Petri dishes and dried at ambient conditions for 24–48 vigorous stirring
h.
for 1 h. The solutions were cast into glass Petri dishes
The resulting PVP/CNC composite films were about 100 µm thick. and dried at ambient conditions for 24–48 h.
The resulting
The aerogels PVP/CNC
based on composite
CNC andfilmsPVPwere
wereabout 100 by
prepared μmfreeze-drying
thick. of aqueous mixtures of the
The aerogels based on CNC and PVP were prepared
two components. The aqueous mixtures of CNC and PVP were prepared by freeze-drying of aqueous mixtures
by ultrasonication of
before the
the two components. The aqueous mixtures of CNC and PVP were prepared by
samples were frozen at −40 ◦ C for 48 h and subsequently lyophilized under reduced pressure of 6 Pa ultrasonication
before
at −54 ◦theC forsamples were frozen at −40 °C for 48 h and subsequently lyophilized under reduced
36–48 h.
pressure of 6 Pa at −54
The composite films °Corfor 36–48 h.
aerogels with different CNC contents (for example, 2 or 4 wt.%, etc.) were
denoted as PVP/CNC-2, PVP/CNC-4, with
The composite films or aerogels different
etc., or CNC contents
PVP/CNC-2 (for example,
aero, PVP/CNC-4 aero,2etc.,
or 4respectively.
wt.%, etc.)
were denoted as PVP/CNC-2, PVP/CNC-4, etc., or PVP/CNC-2 aero, PVP/CNC-4 aero, etc.,
respectively.
2.4. Characterization

2.4.1. Determination of Size and Surface Charge of CNC Particles


2.4. Characterization
The size of CNC particles was determined using an EMV-100L (Sumy, Ukraine) transmission
2.4.1. Determination of Size and Surface Charge of CNC Particles
electron microscope (TEM) (accelerating voltage of 50 kV).
size of CNC
The particle particles wasindetermined
size distribution using
aqueous and an EMV-100L
organic (Sumy, Ukraine)
CNC suspensions transmission
was determined by
electron microscope
dynamic (TEM)
light scattering (accelerating
method voltage of 50
(DLS) (wavelength ofkV).
633 nm) using a Zetasizer Nano ZS (Malvern
The particle
Instruments size distribution
Ltd, Malvern, in aqueous
UK) device. and organic
This equipment CNC suspensions
provides measurement wasof determined by
particles sizes
dynamicfrom
ranging light 0.3
scattering
nm to method
6 µm. The (DLS) (wavelength were
measurements of 633carried
nm) using
out aatZetasizer ◦ C and
NanoofZS20(Malvern
temperature
Instruments Ltd,
concentration Malvern,
of 0.1 mg/mLUK) device. This
in disposable equipment
polystyrene providesThe
cuvettes. measurement of particles
obtained particle sizes
size values
ranging
are from 0.3
the results nm to 6 μm.
of averaging overThe
fivemeasurements were carriedcycles.
consecutive measurement out atThe
temperature of 20 in
value obtained °Ceach
and
concentration
cycle of the
is, in turn, 0.1 mg/mL
result ofinautomatic
disposableprocessing
polystyreneof cuvettes. The obtained particle
10–15 measurements. Measuredsizebyvalues are
the DLS
the results of averaging over five consecutive measurement cycles. The value obtained in each cycle
is, in turn, the result of automatic processing of 10–15 measurements. Measured by the DLS method,
CNC particle sizes are the mean hydrodynamic diameters of equivalent spheres. They do not
represent actual physical dimensions of the rod-like CNC particles but are valid for comparison
Nanomaterials 2018, 8, 1011 4 of 21

method, CNC particle sizes are the mean hydrodynamic diameters of equivalent spheres. They do
not represent actual physical dimensions of the rod-like CNC particles but are valid for comparison
purposes. However, the hydrodynamic diameter strongly correlates with the length of a rod-shaped
particle evaluated by electron microscopy or atomic force microscopy [33,34].
The CNC particle surface charge in aqueous suspension was evaluated by measuring the
ζ-potential (Zetasizer Nano ZS, Malvern Instruments Ltd, Malvern, UK).

2.4.2. Scanning Electron Microscopy (SEM)


The investigations of the surface morphology of the composites were carried out using a VEGA3
TESCAN instrument (Brno, Czech Republic) (accelerating voltage of 5 kV, a chamber pressure of about
9 × 10−3 Pa, contrast in a secondary electron mode).

2.4.3. Fourier Transform Infrared Spectroscopy (FTIR)


The FTIR measurements of the samples (pressed tablets with KBr) were carried out at room
temperature using a VERTEX 80v FTIR spectrometer (Bruker, Ettlingen, Germany) in the range of 4000
to 400 cm−1 .

2.4.4. X-ray Diffraction Analysis


The X-ray diffraction analysis was performed using a Bruker D8 Advance powder diffractometer
(Bruker BioSpin GmbH, Rheinstetten, Germany) according to the Bragg–Brentano scheme with Cu-Kα
radiation (λ = 0.1542 nm).
CNC crystallite sizes were estimated using the Scherrer equation [35]:

L = Kλ/ß1/2 Cosθ, (1)

where, L is the crystal dimension perpendicular to the diffracting planes with hkl Miller indices, λ is the
wavelength of X-ray radiation, ß1/2 is the full width of the diffraction peak measured at half maximum
height (FWHM), θ is the diffraction angle, and K = 0.94.

2.4.5. Thermogravimetric Analysis (TGA)


Thermogravimetric analysis was performed on a TG 209 F1 (Netzsch Gerätebau GmbH,
Selb, Germany) at a heating rate of 10 K min−1 in an atmosphere of dry argon at a flow rate of
30 mL min−1 .

2.4.6. Differential Scanning Calorimetry (DSC)


The DSC measurements were performed on a DSC 204 F1 (Netzsch Gerätebau GmbH,
Selb, Germany) in an atmosphere of ultra pure grade dry argon at a flow rate of 15 mL min−1
and a heating rate of 10 K min−1 using standard aluminum crucibles.

2.4.7. Mechanical Properties


The tensile strength and elongation at break were measured using an IR 5046-5 tensile-testing
machine (Ivanovo, Russia) at a loading rate of 1 mm/min and at room temperature with a solid clamp
in the tension mode [36]. Four specimens with the dimensions of 15 mm (length) × 5 mm (width) ×
0.1 mm (thickness) were used for each sample group. The stress and strain values were calculated
from the machine-recorded force and displacement based on the initial cross-section area and the
original gauge length (10 mm) of each sample, respectively. The Young’s modulus for each sample was
calculated from the initial linear portion of the stress-strain curves through a linear regression analysis.
The obtained values of the Young’s modulus were within ±10%, while the stress and the elongation at
break fluctuated in the range of ±15%. Glycerol was added (2% v/v) to the CNC solution to enhance
the elasticity of the films, and this solution was further used to dissolve PVP.
Nanomaterials 2018, 8, 1011 5 of 21

2.4.8. Water Uptake


Nanomaterials 2018, 8, x FOR PEER REVIEW 5 of 21
Isotherms of water adsorption and desorption upon films and aerogels of the PVP/CNC
composites
25 °C underwere obtained.
controlled The sorption
humidity (aqueousexperiment was
solutions of conducted
sulfuric acid ofinaacertain
desiccator at temperature
concentration). The
of ◦
25 uptake
C under controlled humidity (aqueous solutions of sulfuric acid of a certain concentration).
water was determined gravimetrically.
The water uptake was determined gravimetrically.
3. Results and Discussion
3. Results and Discussion
3.1. FTIR Analysis
3.1. FTIR Analysis
Figure 2 shows FTIR spectra of PVP/CNC composite films. The peaks located at about 1291,
Figure 2 shows FTIR−1spectra of PVP/CNC composite films. The peaks located at about 1291, 1426,
1426, 1666, and 2953 cm were assigned to the stretching vibrations of C–N, C=C, C=O, and C–H of
1666, and 2953 cm−1 were assigned to the stretching vibrations of C–N, C=C, C=O, and C–H of PVP,
PVP, respectively [13]. The appearance of a broad band at about −3440 cm−1 suggests that PVP
1 suggests that PVP contains
respectively [13]. The appearance of a broad band at about 3440 cm
contains adsorbed water. The large bands near 2900 cm−1 and at 3300–3500 cm−1 assigned
adsorbed water. The large bands near 2900 cm−1 and at 3300–3500 cm−1 assigned respectively to
respectively to the C–H and O–H stretching vibrations of cellulose, were found for all the
the C–H and O–H stretching vibrations of cellulose, were found for all the characterized materials.
characterized materials. The characteristic bands at 1060 and 1030 cm−1 and at 1165, corresponded to
The characteristic bands at 1060 and 1030 cm− 1 and at 1165, corresponded to the C–OH stretching
the C–OH stretching vibrations of the cellulose secondary and primary alcohols and to the
vibrations of the cellulose secondary and primary alcohols and to the asymmetric ring breathing mode
asymmetric ring breathing mode of cellulose, respectively [37]. The stretching vibration of the
of cellulose, respectively [37]. The stretching vibration of the cellulose C–O–C bond was identified at
cellulose C–O–C bond was identified at 1110 cm−1 and 1050 cm−1. The FTIR spectra of all the
1110 cm− 1 and 1050 cm− 1 . The FTIR spectra of all the PVP/CNC composites demonstrated an intense
PVP/CNC composites demonstrated − 1
an intense wide band at about 3440 cm−1 related to adsorbed
wide band at about 3440 cm related to adsorbed water. As the CNC content in the composites
water. As the CNC content in the composites increased, almost no changes in the positions of all
increased, almost no changes in the positions of all these peaks were noted. Thus, the FTIR spectra
these peaks were noted. Thus, the FTIR spectra indicated that the molecular interactions between
indicated that the molecular interactions between CNC and PVP are weak (see Section 3.9). However,
CNC and PVP are weak (see Section 3.9). However, the data obtained [38] suggest that in
the data obtained [38] suggest that in complexation reactions with substances of a different chemical
complexation reactions with substances of a different chemical nature, PVP participates in hydrated
nature, PVP participates in hydrated state through bound water. In this case, water can stabilize
state through bound water. In this case, water can stabilize the resulting intermolecular complexes. It
the resulting intermolecular complexes. It is obvious that the PVP interaction with water occurs on
is obvious that the PVP interaction with water occurs on account of hydrogen bonding. It is believed
account of hydrogen bonding. It is believed that a PVP unit forms hydrogen bonds with two water
that a PVP unit forms hydrogen bonds with two water molecules through lone-electron pairs of the
molecules through lone-electron pairs of the oxygen atom of the carbonyl group of the pyrrolidone
oxygen atom of the carbonyl group of the pyrrolidone heterocycle [38].
heterocycle [38].

1
0 1 0
2 2
3
4 50 3
50 5 4
Transmittance, %

Transmittance, %

6
7 100 5
6
100
8 150 7
8
200
150

250

200
4000 3500 3000 2500 2000 1500 1000 500
-1 -1
Wavenumber, cm Wavenumber, cm

(a) (b)
Figure 2. FTIR
FTIRspectra
spectraofofthethePVP/CNC
PVP/CNC composite
compositefilms: 1 - neat
films: 1—neatPVP;PVP;
2 - PVP/CNC-4.6;
2—PVP/CNC-4.6;3 -
PVP/CNC-10.9; 4 - PVP/CNC-16.3; 5 - PVP/CNC-19.6; 6 - PVP/CNC-28.9;
3—PVP/CNC-10.9; 4—PVP/CNC-16.3; 5—PVP/CNC-19.6; 6—PVP/CNC-28.9; 7—PVP/CNC-37.9; 7 - PVP/CNC-37.9; 8 - neat
CNC: (a) in the range of wavenumbers of 4000–2500 −1; −
cmcm (b) in in
thethe
range
8—neat CNC: (a) in the range of wavenumbers of 4000–2500 1 ; (b) rangeofofwavenumbers of
wavenumbers of
2500–500 cm
2500–500 cm . .PVP,
−1
− 1 PVP,polyvinylpyrrolidone;
polyvinylpyrrolidone;CNC,CNC,cellulose
cellulosenanocrystals.
nanocrystals.

3.2. Water Adsorption upon the PVP/CNC Composites


Water sorption
sorption is one of the major
major problems limiting applications for CNC based materials.
materials.
Water
Water uptake
uptake affects
affects mechanical
mechanical properties,
properties, chemical
chemical and
and dimensional stabilities. On the other hand,
the water absorption capacity together with the biodegradability are among the most important
important
properties for the practical applications and post-use of the biodegradable materials.
the practical applications and post-use of the biodegradable materials. There is a
plethora of studies that have investigated the moisture sorption and retention by CNC reinforced
polymer composites [39–44]. Absorbed water molecules can be accumulated in three regions within
a CNC-based nanocomposite: on the CNC, in a polymer matrix, and at the polymer/CNC interface
[45]. The accessible hydroxyl groups on the CNC particles surface are commonly considered as the
Nanomaterials 2018, 8, 1011 6 of 21

plethora of studies that have investigated the moisture sorption and retention by CNC reinforced
polymer composites [39–44]. Absorbed water molecules can be accumulated in three regions within a
Nanomaterials 2018, 8, x FOR PEER REVIEW 6 of 21
CNC-based nanocomposite: on the CNC, in a polymer matrix, and at the polymer/CNC interface [45].
The accessible hydroxyl groups on the CNC particles surface are commonly considered as the initial
initial site of water sorption. However, for polymer composites, the CNC networks can decrease
site of water sorption. However, for polymer composites, the CNC networks can decrease water
water sorption as compared to the neat polymer matrices that readily adsorb water [46].
sorption as compared to the neat polymer matrices that readily adsorb water [46].
Isotherms of water adsorption and desorption upon films and aerogels of the PVP/CNC
Isotherms of water adsorption and desorption upon films and aerogels of the PVP/CNC
composites were obtained. The adsorption experiment was carried out as described earlier [32]. A
composites were obtained. The adsorption experiment was carried out as described earlier [32].
detailed analysis of the isotherms is depicted in Figure 3. Obviously, water uptake increases with a
A detailed analysis of the isotherms is depicted in Figure 3. Obviously, water uptake increases with a
relative pressure increase. Meanwhile, the water uptake on the composites decreases gradually with
relative pressure increase. Meanwhile, the water uptake on the composites decreases gradually with
the CNC content increase. Water uptake by the aerogels is very similar to the uptake by the films in
the CNC content increase. Water uptake by the aerogels is very similar to the uptake by the films in
spite of their different densities (see Section 3.3).
spite of their different densities (see Section 3.3).

100
90
90
80
P/P0=1 80
70 P/P0=1
70
Water uptake, %

Water uptake, %
60
P/P0=0.959 60 P/P0=0.959
50 P/P0=0.883
P/P0=0.883 50
40
40
30 P/P0=0.754 P/P0=0.754
30
20
P/P0=0.579 20 P/P0=0.579
P/P0=0/358 P/P0=0.358
10
P/P0=0.166 10 P/P0=0.166
0 P/P0=0.037 0
P/P0=0.037
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40

CNC content, % CNC content, %

(a) (b)
100 100

90 90

80 80
Water uptake, %

70 70
Water uptake, %

P/P0=1
60 60 P/P0=1
50 50

40
P/P0=0.959 40 P/P0=0.959
30 30 P/P0=0.883
P/P0=0.883 P/P0=0.754
20
20 P/P0=0.754
P/P =0.579 P/P0=0.579
10 P/P00=0.358 10 P/P =0.358
P/P0=0.166 P/P00=0.166
0 0 P/P0=0.037
0 10 20 30 40 50 60 70
P/P0=0.037
80 0 10 20 30 40 50 60 70

CNC content, % CNC content, %

(c) (d)
Figure 3.3.Equilibrium
Equilibriumwater content
water in the
content in composite films films
the composite (a,b) and
(a, aerogels (c,d) at water
b) and aerogels (c, d)adsorption
at water
(a,c) and desorption (b,d) depending on the CNC content (P/P is a relative pressure of water
adsorption (a, c) and desorption (b, d) depending on the CNC0 content (P/P0 is a relative pressure vapor).
of
water vapor).
As PVP is a hydrophilic polymer, it has a great tendency to adsorb water. Evidently, the water
uptake
Asby
PVPtheiscomposites
a hydrophilicdoes not depend
polymer, it hason eithertendency
a great morphology (film or
to adsorb aerogel)
water. or density
Evidently, the of the
water
sample, but is determined only by the PVP content in the composite. The molecular
uptake by the composites does not depend on either morphology (film or aerogel) or density of theinteractions
between CNC
sample, but and PVP areonly
is determined weak,
by but they content
the PVP may interact
in thewith each other
composite. The through
molecularbound water.
interactions
Apparently,
between CNC theand
reduced
PVP water binding
are weak, but capacity
they mayofinteract
the PVP/CNC
with eachcomposites as compared
other through to the
bound water.
neat PVP can be related to the increase in the intramolecular CNC–CNC and PVP–PVP
Apparently, the reduced water binding capacity of the PVP/CNC composites as compared to the interactions
(see
neatSections
PVP can3.5beand 3.9).to the increase in the intramolecular CNC–CNC and PVP–PVP interactions
related
(see Sections 3.5, 3.9).
3.3. Density of the PVP/CNC Composite Films
3.3. Density
Figure 4ofshows
the PVP/CNC Composite
the composites Films versus the CNC content.
density
Figure 4 shows the composites density versus the CNC content.
Nanomaterials
Nanomaterials2018, 8, 8,
2018, 1011
x FOR PEER REVIEW 77ofof2121

1,8
0,06
1,7

1,6 0,05
-3

-3
Density, g cm 1,5

Density, g cm
0,04
1,4

1,3 0,03

1,2
0,02
1,1

1,0 0,01

0,9
0,00
0 20 40 60 80 100 0 10 20 30 40 50 60 70 80
CNC content, % CNC content, %

(a) (b)

Figure Density
4. 4.
Figure DensityofofPVP/CNC
PVP/CNCcomposites
compositesasasa afunction
functionofofthe
theCNC
CNCcontent:
content:(a)
(a)films;
films;(b)
(b)aerogels.
aerogels.
The error
The bars
error areare
bars thethe
standard deviations
standard in in
deviations repeated experiments.
repeated experiments.

The
Thedensity
density waswas determined
determined by by
measuring
measuring the the
samples geometry
samples geometry and anddividing the weight
dividing by theby
the weight
volume. The reported results are the average values obtained from at
the volume. The reported results are the average values obtained from at least three different least three different samples.
It samples.
can be seen that be
It can theseen
dependence
that theof the film density
dependence of the onfilm
the CNC
densitycontent
on the is of
CNCcomplex nature:
content is of the film
complex
density is maximal at the CNC content of about 5–10 wt% and minimal
nature: the film density is maximal at the CNC content of about 5–10 wt% and minimal at the CNC at the CNC content of about
55–60 wt%.
content of However,
about 55–60 thewt%.
aerogel density gradually
However, the aerogel decreases with the increase
density gradually decreasesin the
withCNCthecontent
increaseinin
thethecomposite.
CNC content in the composite.
The
Thestrange
strangedensity
densitybehavior
behaviorofofthe thecomposite
compositefilms filmsdeserves
deservesa aspecialspecialattention
attentionand andsome
some
speculation.
speculation. If we assume that in the composite films the PVP molecules are in a coil (orglobule)
If we assume that in the composite films the PVP molecules are in a coil (or globule)
conformation,
conformation, then
thentheir shape
their can be
shape canconsidered
be consideredto be spherical (see Section
to be spherical (see 3.9). The phase
Section 3.9). Thediagram
phase
for packing of mixtures of spheres and rods is very complex because it
diagram for packing of mixtures of spheres and rods is very complex because it depends on the depends on the stoichiometry,
the rod aspect ratio,
stoichiometry, andaspect
the rod size ratio ofand
ratio, the rod
size and
ratiothe
of sphere
the rod(R and the/R
sphere rod ). Simulations
sphere (Rsphere/Rrod).predict the
Simulations
densest packing for binary structure of spheres and rods (of 2:1 stoichiometry)
predict the densest packing for binary structure of spheres and rods (of 2:1 stoichiometry) at an at an appropriate
Rsphere /Rrod range,
appropriate Rspherethat
/Rrodexceeds
range, thethatdensities
exceeds of thedensities
the demixedofphases (rods and
the demixed spheres)
phases [47].
(rods andHowever,
spheres)
at[47].
intermediate
However, volume fractions, simulations
at intermediate predictsimulations
volume fractions, phase separation
predict into rodsseparation
phase and spheres. Moreover,
into rods and
depending
spheres. on volume fractions,
Moreover, depending size on
ratiovolume
or aspectfractions,
ratio, packing
size ofratio
phase-separated
or aspect ratio,rods and spheresof
packing
may be more effective than packing of binary structure of spheres and
phase-separated rods and spheres may be more effective than packing of binary structure of spheres rods. Thus, we observe a
complex density
and rods. Thus,behavior
we observe of thea composite films asbehavior
complex density a result ofofthe
thepacking
composite efficiency
films ofas rod-like
a result CNCof the
particles and spherical PVP molecules.
packing efficiency of rod-like CNC particles and spherical PVP molecules.

3.4.
3.4.TGTGAnalysis
Analysis
The
TheTG
TGand
and derived
deriveddifferential TG
differential (DTG)
TG curves
(DTG) curvesofof
thethe
PVP/CNC
PVP/CNCcomposite
compositefilms
filmsare
areshown
shown
ininFigure S2a,b, respectively. The thermal parameters, including the maximum thermal degradation
Figure S2a,b, respectively. The thermal parameters, including the maximum thermal degradation
temperature
temperature (T(T
max ) and the onset thermal degradation temperature (Ton ) are summarized in Table 1.
max) and the onset thermal degradation temperature (Ton) are summarized in Table 1.

Table 1. Results of TG analysis of the PVP/CNC composite films.


Table 1. Results of TG analysis of the PVP/CNC composite films.
Initial Mass Loss
Initial mass loss
FirstFirst
Degradation Stage
degradation stage
Second Degradation Stage
Second degradation stage Total
Sample Total mass
Sample T on , T max , Mass T on ,
Mass T max , MassMass T on , T max , Mass Mass Mass
Ton
◦ ,C°C T◦max
C , °C Loss, %% ◦ T
Con, °C ◦ Tmax, °C
C Loss, loss,
% % C ◦ T on, °C ◦ Tmax, °C
C Loss, % Loss, loss,
% %
loss, loss, %
PVP
PVP - 76.4
76.4 14.614.6 - - - - - - 405.4405.4432.3 432.3 80.9 80.9 95.5 95.5
PVP/CNC-2.3
PVP/СNС-2.3 - 77.9
77.9 15.115.1 - - - - - - 403.3403.3432.6 432.6 78.4 78.4 93.5 93.5
PVP/CNC-6.7
PVP/СNС-6.7 - 80.6
80.6 13.713.7 240.1
240.1 264.5264.5 3.9 3.9 402.9402.9433.8 433.8 76.6 76.6 94.2 94.2
PVP/CNC-10.9
PVP/СNС-10.9 - 69.0
69.0 12.812.8 239.5
239.5 259.8259.8 6.2 6.2 402.9402.9434.9 434.9 73.7 73.7 92.7 92.7
PVP/CNC-19.6
PVP/СNС-19.6 -- 72.8
72.8 12.212.2 232.6
232.6 252.1252.1 11.7 11.7 403.3403.3436.6 436.6 64.0 64.0 87.9 87.9
PVP/CNC-28.9 - 66.6 11.0 229.8 246.3 16.0 401.9 434.1 59.9 86.9
PVP/СNС-28.9
PVP/CNC-37.9 -- 66.6
64.8 10.611.0 229.8 242.4246.3
226.4 19.6 16.0 398.2401.9430.4 434.1 53.8 59.9 84.0 86.9
PVP/СNС-37.9
CNC -
63.0 64.8
82.2 3.010.6 226.4 169.8242.4
154.5 37.8 19.6 334.1398.2376.4 430.4 30.1 53.8 70.9 84.0
CNC 63.0 82.2 3.0 154.5 169.8 37.8 334.1 376.4 30.1 70.9
Nanomaterials 2018, 8, 1011 8 of 21

The PVP-based composites with a small CNC content have one main characteristic mass loss
region (excluding the initial region of water desorption). On the thermograms of the composites with
the CNC content of about 7% or more, two main mass loss regions are observed. The water content in
the samples gradually decreases with the growing CNC content, and for the neat CNC it equals 3.0%.
The first main degradation range is located between ~230 and 400 ◦ C and is attributed to pyrolysis
of the CNC catalyzed by acid sulfate groups [48]. The second stage mass loss occurs at about 400 ◦ C
and involves decomposition of PVP and CNC carbonaceous matter [31].

3.5. DSC Analysis


Before the DSC scanning, all the samples were first equilibrated at 20 ◦ C and heated to 200 ◦ C at
a heating rate of 10 K min−1 , and then cooled down to 20 ◦ C at the same rate (Figure S3). The glass
transition temperatures (Tg) for all the samples determined under heating and cooling are reported
in Table 2. The results of DSC analysis of the PVP/CNC films show that Tg tends to shift to higher
values with increasing the CNC content in the composite. This higher Tg value can point out that
the association of PVP molecules is enhanced by the CNC presence. A notable increase in Tg may be
ascribed to the macromolecular confinement provided by the CNC surfaces [49]. On the other hand,
the nanoconfinement effects may facilitate significant enhancement of mechanical properties.

Table 2. Results of DSC analysis of the PVP/CNC composite films.

Tg, ◦C
Sample
Heating Cooling
PVP 180.2 166.7
PVP/CNC-4.6 176.5 173.9
PVP/CNC-10.6 178.3 174.7
PVP/CNC-19.9 184.2 177.7
PVP/CNC-28.9 198.5 186.4
PVP/CNC-37.9 198.6 184.1

However, analyzing the data on density, as well as the results of TG and DSC analysis, one can
observe an interesting feature. Composite films with a CNC content of about 5–10% are characterized
by the maximum density, the maximum temperature of water desorption and the minimum Tg under
heating. Apparently, this is due to the firmly bound water, which is not completely removed from PVP
even under heating up to 250 ◦ C [38]. Indeed, the strongly bound water can act as a plasticizer and
increase mobility of the polymer chains (the Tg decreases).

3.6. Mechanical Properties


Figure S4 shows stress-strain curves of the PVP/CNC composite films. The values of ultimate
tensile strength (σmax ), Young’s modulus (E), and elongation at break (εb ) are summarized in Table 3.

Table 3. Tensile properties of the PVP/CNC composite films.

Ultimate Tensile Elongation at Young’s Modulus


Sample
Strength (σ max ), MPa Break (εb ), % (E), MPa
PVP 6.6 5.5 120
PVP/CNC-4.4 9.7 7.4 132
PVP/CNC-8.4 17.2 4.2 413
PVP/CNC-17.7 31.8 6.9 457
PVP/CNC-26.8 38.4 3.0 1284
PVP/CNC-29.2 43.8 2.5 1755
PVP/CNC-39.9 25.3 2.1 1216
Nanomaterials 2018, 8, 1011 9 of 21

Upon the addition of CNC, σmax of the composites increases, indicating a reinforcing effect of
the CNC. However, when CNC is added, εb decreases gradually, which suggests that the composites
become brittle compared with the neat PVP. As the CNC content grows, σmax and E increase and reach
maxima at the CNC content of 29.2 wt.%. However, upon further increasing of the CNC content, σmax
and E decrease. Therefore, a CNC content of about 30 wt.% is optimal for the composite films based on
the σmax and E values.
Various modeling approaches have investigated the effect of CNC fillers within polymer matrices
on effective nanocomposite properties. Typically, Halpin–Tsai, Halpin–Kardos, and percolation
approaches have been used to understand the reinforcing effect of low concentration CNC in
low-modulus polymers [2]. Experimentally obtained tensile modulus could be compared with
theoretical predictions by using a modified Halpin–Tsai model as [50]:

Ec /Em = (3/8)·(1 + 2ρη L V CNC )/(1 − η L V CNC ) + (5/8)·(1 + 2η T V CNC )/(1 − η T V CNC ),
η L = (Er − 1)/(Er + 2ρ), (2)
η T = (Er − 1)/(Er + 2),

where Ec and Em are the Young’s modulus of the composite and matrix, respectively; ρ is the CNC
aspect ratio; Er is the ratio between the Young’s modulus of CNC and the matrix; V CNC is CNC volume
fraction in the composite.
Based on the weight percentage (WCNC ), one can estimate the corresponding volume fraction
V CNC by knowing the density of CNC and PVP as:

V CNC = (W CNC /dCNC )/(W CNC /dCNC + (1 − W CNC )/dPVP ) (3)

The Young’s modulus of CNC nanoparticles has been set at 105 GPa; the CNC aspect ratio has been
assumed as 7; the densities for CNC and PVP have been assumed as 1.46 and 1.2 g/cm3 , respectively.
Figure 5 compares the experimentally obtained data and predicted Young’s modulus values
according to random Halpin–Tsai equation. The Halpin–Tsai model tends to lower predictions for
high modulus ratio systems Er , that is common in nanocomposites based on rubbery matrices [2].
These predicted values are not accurate for composites with aligned stiff short fillers. Besides inaccuracy
at high modulus ratio, the Halpin–Tsai equation has no information about filler anisotropy or about
the quality of the interface. However, a stiffening of the interphase region might be very important and
useful for stress transfer from the matrix to the CNC. Thus, an observation that experiments exceed
the Halpin–Tsai model is the expected result in case of the CNC filler and a low-modulus polymer.

Figure 5. Experimental data (1) and fitting results according to Halpin–Tsai model (2). The error bars
are the standard deviations in repeated experiments.
Nanomaterials
Nanomaterials 8, x8,FOR
2018,
2018, 1011PEER REVIEW 10 21
10 of of 21

The observed experimental results exceeding Halpin–Tsai model can be explained in terms of
The observed
percolation concept in experimental
CNC basedresults exceeding
composites. TheHalpin–Tsai
concept of model can be
percolation is explained
related with in terms
the
of percolation concept in CNC based composites. The concept of percolation
development of a connected network in a multiphase system. The percolating phase reaches the is related with the
development
percolation of a connected
threshold, i.e., a network
connectedin network
a multiphase system. when
is formed, The percolating phase reaches
the concentration of the the
percolation threshold, i.e., a connected network is formed, when the concentration
percolating phase increases. Some effective properties increase rapidly and dramatically for of the percolating
phase increases.
concentrations Some
above theeffective
percolationproperties increase
threshold. rapidly and dramatically
In nanocomposites, percolationforisconcentrations
most dramaticabove for
the percolation
high modulus ratiothreshold.
systems,In nanocomposites,
and the percolationpercolation is mostcan
threshold dramatic for hightomodulus
be shifted very low ratio
systems, and the percolation threshold can be
concentrations by using high aspect ratio nanofillers [51].shifted to very low concentrations by using high aspect
ratio
Thenanofillers [51].
Young’s modulus and tensile strength of CNC reinforced polymer matrix composites can
The Young’s
be compared to other modulus and tensile
material systemsstrength of CNC
via ‘Ashby reinforced
plots’. polymer
According to matrix composites
Gibson–Ashby can be
theory,
compared to other material systems via ‘Ashby plots’. According to Gibson–Ashby
Young's modulus and strength of a solid are proportional to the relative density [52]. Possible theory, Young’s
modulus
reasons for and
poorstrength
enough of a solid areproperties
mechanical proportional to thebased
in CNC relative density [52].
composites Possible
are CNC reasons for
aggregation,
low interfacial properties of CNC/matrix, low particle aspect ratio, and low CNC alignmentinterfacial
poor enough mechanical properties in CNC based composites are CNC aggregation, low [2]. The
properties of CNC/matrix, low particle aspect ratio, and low CNC alignment
experimental data obtained are in good agreement with the density data, according to which the [2]. The experimental
data obtained
density are in films
of composite good decreases
agreementsharply
with thewith
density
a CNCdata, according
content to which
of more thanthe
30% density
(Figureof 4a).
composite
films decreases sharply with a CNC content of more than 30% (Figure 4a).
3.7. X-ray Diffraction Analysis
3.7. X-ray Diffraction Analysis
The
TheX-ray diffraction
X-ray patterns
diffraction ofof
patterns the PVP/CNC
the films
PVP/CNC areare
films shown in in
shown Figure 6. 6.
Figure

PVP
2000
(200) PVP/CNC-4.6
PVP/CNC-10.9
PVP/CNC-19.6
PVP/CNC-28.9
PVP/CNC-37.9
PVP/CNC-54.5
(1-10) PVP/CNC-70.6
(110)
Intensity

1000

0
5 10 15 20 25 30 35

2θ, deg.

Figure 6. X-ray diffraction patterns of the PVP/CNC composite films.


Figure 6. X-ray diffraction patterns of the PVP/CNC composite films.
As PVP is an amorphous polymer, the diffraction peak arising at about 2θ = 22.9 is attributed to the
As PVP is an amorphous polymer, the diffraction peak arising at about 2θ = 22.9 is attributed to
(200) plane of cellulose Iß , whereas the two overlapped weaker diffractions at 2θ close to 16.6◦ and 14.8◦
the (200) plane of cellulose Iß, whereas the two overlapped weaker diffractions at 2θ close to 16.6°
are assigned to the (110) and (1–10) lattice planes of cellulose Iß [53]. The intensity of the diffraction
and 14.8° are assigned to the (110) and (1–10) lattice planes of cellulose Iß [53]. The intensity of the
peaks for CNC becomes more pronounced when the CNC content in the composite is increased.
diffraction peaks for CNC becomes more pronounced when the CNC content in the composite is
The powder diffraction peaks were deconvoluated with the Pseudo-Voigt function, assuming a
increased.
linear background [54]. The average dimensions of the elementary crystallites in the diffracting planes
The powder diffraction peaks were deconvoluated with the Pseudo-Voigt function, assuming a
were evaluated by applying the Scherrer equation (1). From these values, we suppose the crystallites
linear background [54]. The average dimensions of the elementary crystallites in the diffracting
have a square cross-sectional geometry with widths of about 2.6 nm and a diagonal of 3.9 nm (Table 4,
planes were evaluated by applying the Scherrer equation (1). From these values, we suppose the
Figure 7).
crystallites have a square cross-sectional geometry with widths of about 2.6 nm and a diagonal of 3.9
nm (Table 4, Figure 7).Table 4. CNC crystallite sizes (nm) in the PVP/CNC composite films.

Table 4. CNC crystallite


Sample sizes (nm)Crystallographic
in the PVP/CNC Plane
composite films.
(1–10) (110) (200)
Crystallographic plane
Sample PVP/CNC-28.9 2.96 2.62 3.74
(1–10) (110) (200)
PVP/CNC-37.9 2.15 2.53 4.19
PVP/CNC-28.9 PVP/CNC-54.5
2.96 2.67 2.62
2.37 3.81 3.74
PVP/CNC-37.9 PVP/CNC-70.6
2.15 2.48 2.53
2.42 3.74 4.19
PVP/CNC-54.5 2.67 2.37 3.81
PVP/CNC-70.6 2.48 2.42 3.74
Nanomaterials 2018,
Nanomaterials 2018, 8, x FOR PEER REVIEW 11 of 21
Nanomaterials 2018, 8,8,x1011
FOR PEER REVIEW 1111ofof21
21

Figure 7.
Figure 7. Cross-section
Cross-section ofofelementary
elementarycrystallites
crystallitesininCNC
CNCparticles.
particles.Projection
Projectionofof cellulose
cellulose molecules
molecules is
Figure 7. Cross-section of elementary crystallites in CNC particles. Projection of cellulose molecules
is clearly visible. The indexation of corresponding lattice planes is described according
clearly visible. The indexation of corresponding lattice planes is described according to the monoclinicto the
is clearly visible. The indexation of corresponding lattice planes is described according to the
monoclinic
unit unit cell ofIallomorph
cell of allomorph ß [55]. Iß [55].
monoclinic unit cell of allomorph Iß [55].
3.8. Morphologies of
3.8. Morphologies of the
the PVP/CNC
PVP/CNC Composites
Composites
3.8. Morphologies of the PVP/CNC Composites
SEM
SEM images
images were
were recorded
recorded to
to analyze
analyze the
the microstructure
microstructure of
of the
the PVP/CNC composite films
PVP/CNC composite films and
and
SEM (Figure
aerogels images 8,
were recorded
Figure 9 and to analyze
Figure S5). the microstructure of the PVP/CNC composite films and
aerogels (Figures 8, 9 and S5).
aerogels (Figures 8, 9 and S5).

(a) (b)
(a) (b)

(c) (d)
(c) (d)

(e) (f)
(e) (f)
8. Cross-sectional
Figure 8. Cross-sectionalSEMSEMimages
imagesofof thethe
neat PVP
neat PVPfilmfilm
(a) and the PVP/CNC
(a) and the PVP/CNCcomposite filmsfilms
composite with
Figure 8. Cross-sectional SEM images of the neat PVP film (a) and the PVP/CNC composite films
CNC content
with CNC (wt.%)
content of: 4.6of:
(wt.%) (b);
4.616.3
(b);(c);
16.328.9
(c);(d);
28.954.5
(d);(e); 85.7
54.5 (e);(f), respectively.
85.7 The scale
(f), respectively. Thebar is 2bar
scale µm.is 2
with CNC content (wt.%) of: 4.6 (b); 16.3 (c); 28.9 (d); 54.5 (e); 85.7 (f), respectively. The scale bar is 2
μm.
μm.
Nanomaterials 2018, 8, 1011 12 of 21
Nanomaterials 2018, 8, x FOR PEER REVIEW 12 of 21

(a) (b)

(c) (d)

(e) (f)
Figure 9. SEM images of the neat
neat PVP
PVP aerogel
aerogel (a)
(a) and
and the
thePVP/CNC
PVP/CNC composite aerogels with CNC
content (wt.%) of: 4.6 (b); 10.9 (c); 19.6 (d); 28.9 (e); 70.6 (f), (f),
content (wt.%) of: 4.6 (b); 10.9 (c); 19.6 (d); 28.9 (e); 70.6 respectively.
respectively. Scales:
Scales: 2 μm2 (a,
µmc,(a,c,d,e,f);
d, e, f); 1
1μm (b).
µm(b).

The cross-sectional
The cross-sectional SEM
SEM image
image ofof the
the neat
neat PVP
PVP film
film showed
showed aa smooth
smooth morphology
morphology (Figure
(Figure 8a).
8a).
With
With the
the increase
increase in
in the
the CNC
CNC content
content in in the
the composite
composite films,
films, the
the fractured
fractured surface
surface became rougher
became rougher
due to the CNC particles aggregation (Figure 8b–f). Separate CNC particles
due to the CNC particles aggregation (Figure 8b–f). Separate CNC particles in the films were notin the films were not
observed. Analyzing
observed. Analyzing SEMSEMimages
imagesof ofthethe
composite
compositeaerogels, we unexpectedly
aerogels, we unexpectedly foundfound
that inthat
a certain
in a
composition range (the CNC content from about 5 to 30 wt.%), the individual CNC
certain composition range (the CNC content from about 5 to 30 wt.%), the individual CNC particles particles became
visible (Figure
became visible 9b–e).
(Figure 9b–e).
Using the
Using the SEM
SEM images,
images, we
we estimated
estimated the the CNC
CNC particle size in
particle size in the
the composite
composite aerogels. For this
aerogels. For this
purpose, we measured the widths and lengths of a set of 100 particles manually
purpose, we measured the widths and lengths of a set of 100 particles manually (Figures S6–S9). (Figures S6–S9).
The analysis
The analysis results
results are
are shown
shown as as histograms
histograms in in Figure
Figure 10.
10.
Nanomaterials 2018, 8, 1011 13 of 21
Nanomaterials 2018, 8, x FOR PEER REVIEW 13 of 21

60
30
PVP/CNC-4.6 aero
PVP/CNC-4.6 aero PVP/CNC-10.5 aero
PVP/CNC-10.5 aero PVP/CNC-19.6 aero
25
PVP/CNC-19.6 aero PVP/CNC-28.9 aero

Frequency, %
PVP/CNC-28.9 aero
Frequency, %

40
20

15

20
10

0
0
20 30 40 50 60 70 80 90 100 110 120
0 500 1000 1500
Particle width, nm Particle length, nm

(a) (b)
Figure
Figure10.
10.The
The CNC particle size
CNC particle sizedistribution
distributionininthe
the composite
composite aerogels
aerogels across
across the widths
the widths (a) along
(a) and and
the lengths
along (b). (b).
the lengths

Theanalysis
The analysisofofthe theimages
imagesshows
showsthat thatthe
thewidths
widthsofofthe theCNC
CNCparticles
particlesininthe thecomposite
compositeaerogels
aerogels
are about 55–65 nm and do not depend on the composition.
are about 55–65 nm and do not depend on the composition. Meanwhile, the lengths of the CNCMeanwhile, the lengths of the CNC
particles in
particles in the
theaerogels
aerogelsareare much
much larger thanthan
larger the lengths of the CNC
the lengths of the particles in the initial
CNC particles in suspension,
the initial
these composites were produced from. In addition, as the CNC content
suspension, these composites were produced from. In addition, as the CNC content in the composite in the composite grows,
the particles
grows, lengthslengths
the particles increase gradually
increase and exceed
gradually 1 micron
and exceed for the for
1 micron composition
the compositionwith a CNC
with acontent
CNC
of 28.9 wt.%. The fraction of particles with the lengths exceeding 600
content of 28.9 wt.%. The fraction of particles with the lengths exceeding 600 nm also increases. nm also increases.
ItIt is
is reasonable
reasonabletotoassume assume thatthat
suchsuchan increase
an increasein the in
CNC theparticles size is due
CNC particles to their
size aggregation.
is due to their
It is well known
aggregation. It isthat
wellCNCknownstrongly
that agglomerate
CNC strongly to stack up in parallel
agglomerate to stackgivingup an in apparent increaseanin
parallel giving
particle width
apparent increase[54].inHowever, it should
particle width be However,
[54]. noted that in this casebetheir
it should notedlengths
that inincrease, while
this case thelengths
their widths
remain unchanged.
increase, while the widths Evidently, PVPunchanged.
remain assists CNCEvidently,
self-assembly PVPwhich
assists produces reasonably uniform
CNC self-assembly which
CNC aggregates
produces reasonably with a high CNC
uniform aspectaggregates
ratio (length/width).
with a high The aspect
aspect ratioratio of the CNCThe
(length/width). aggregates
aspect
increase
ratio of the as CNC
the CNC content increase
aggregates grows from as theabout
CNC 5 tocontent
30 wt.% because
grows fromtheabout
lengths5 toincrease
30 wt.%up to fourfold,
because the
while the
lengths widths
increase upremain constant.
to fourfold, while Apparently,
the widths here remainwe constant.
observe the result of ahere
Apparently, synergistic
we observeeffect
theof
PVP adsorption
result onto CNC
of a synergistic particles
effect of PVPand ice-templating
adsorption onto CNCunderparticles
freeze-drying [56,57]. As inter-chain
and ice-templating under
hydrogen bonding is prevailing within the (200)
freeze-drying [56,57]. As inter-chain hydrogen bonding is prevailingplane for cellulose Iß (the so-called
within the ”hydrogen-bonded”
(200) plane for
plane) [2,58,59],
cellulose one can suppose
Iß (the so-called that adsorbed
”hydrogen-bonded” PVP [2,58,59],
plane) blocks theone possibility
can suppose of forming lateral bonds
that adsorbed PVP
between the CNC particles and, thus, prevents their aggregation. Freezing
blocks the possibility of forming lateral bonds between the CNC particles and, thus, prevents their of CNC suspensions can
align rod-like
aggregation. particles
Freezing of because the particles
CNC suspensions cancan be subjected
align to an elongation
rod-like particles because the action as thecan
particles speed
be
of ice crystals
subjected to angrowth
elongationalong the length
action as thediffers
speed fromof icetheir growth
crystals across
growth the width
along [60,61].
the length Moreover,
differs from
somegrowth
their tribology findings
across highlight
the width that Moreover,
[60,61]. polymer adsorption on CNC
some tribology surfacehighlight
findings together that
with polymer
the CNC
surface hydration
adsorption on CNCcan act as together
surface a lubrication
withand thereduce a friction
CNC surface in the CNC
hydration cancontacts
act as aconsiderably
lubrication and [46].
reduce a friction in the CNC contacts considerably [46].
3.9. PVP Adsorption onto CNC Particles
3.9. PVPPVPAdsorption
as a linear ontoamorphous
CNC Particles polymer with a large heterocyclic unit in the polymer chain is
characterized
PVP as a by a coil-globule
linear amorphous conformational
polymer with transition
a largeinheterocyclic
solution, which
unitisincaused by the interaction
the polymer chain is
of distant units in the chain. If the process is driven by units repulsion,
characterized by a coil-globule conformational transition in solution, which is caused the coil swells. If the monomer
by the
units are attracted
interaction of distantto each
unitsother,
in thethe polymer
chain. If the chain condenses
process onto
is driven byitself into
units a dense conformation—the
repulsion, the coil swells. If
so-called
the monomer polymer globule.
units are It isto
attracted thought that the
each other, at higher
polymer temperatures or in aonto
chain condenses good solvent,
itself into amutual
dense
conformation—the so-called polymer globule. It is thought that at higher temperatures or in aAt
repulsion of the polymer units prevails, and the coil swells due to the excluded volume effect. low
good
temperatures
solvent, mutual or repulsion
in a poor solvent, the chains
of the polymer are mostly
units prevails,attracted
and thetocoil
eachswells
other,dueand to
thethecoilexcluded
collapses
into a globule. A very small attraction between the polymer chain units
volume effect. At low temperatures or in a poor solvent, the chains are mostly attracted to is usually enough to cause
each
the polymer to condense into a globule. In polymers with flexible chains (as
other, and the coil collapses into a globule. A very small attraction between the polymer chain units PVP), the coil-globule
istransition is a continuous
usually enough to causeprocess.
the polymer to condense into a globule. In polymers with flexible chains
Figure 11 shows
(as PVP), the coil-globule the transition
size distribution of PVP particles
is a continuous process.depending on the concentration in solution
and Figure
their surface charge determined by the ζ-potential
11 shows the size distribution of PVP particles value. With an increaseonin the
depending the PVP concentration
concentration in
from 0.05 to 7%, the fraction of particles with the size of 60–70 nm disappears,
solution and their surface charge determined by the ζ-potential value. With an increase in the PVP while the fraction
concentration from 0.05 to 7%, the fraction of particles with the size of 60–70 nm disappears, while
Nanomaterials
Nanomaterials2018,
2018,8,8,x1011
FOR PEER REVIEW 1414
ofof2121
Nanomaterials 2018, 8, x FOR PEER REVIEW 14 of 21
the fraction of particles with sizes less than 10 nm grows. At increased PVP concentrations, the
of particles
the
average size with
fraction of sizes
of these
particlesless than
with
particles 10
issizes nm
lessgrows.
reduced than At nm
10 increased
approximately grows.
to 3–6 PVP
Atnm.concentrations,
increased the average
PVP concentrations,
The experimental values ofsize of
the
the
these
averageparticles
size of is reduced
these approximately
particles is reduced to 3–6 nm.
approximately The experimental
to 3–6 nm. values
The
PVP particle size obtained by dynamic light scattering are in good agreement with the estimated of the
experimental PVP particle
values of size
the
obtained
PVP
ones. Thus,bythe
particle dynamic
size
length light
obtained
of thescattering
by dynamic
expanded arepolymer
in good
light agreement
scattering
chain ofarePVP with thethe
inwith
good estimated
agreement
molecularones.
withThus,
massthe the length
of estimated
40,000 is
of the expanded polymer chain of PVP with the molecular mass of 40,000 is
about 50 nm, and the radius of gyration of the coil, Rg, varies from 1 to 7 nm that can be calculatedis
ones. Thus, the length of the expanded polymer chain of PVP with the molecularabout 50
mass nm,
of and
40,000 the
radius
about
from ofknown
50
the gyration
nm, and of the
the coil, R
radius
relations ofg ,gyration
varies from 1 tocoil,
of the 7 nm Rgthat can be
, varies fromcalculated
1 to 7 nmfrom the
that known
can relations
be calculated
from the known relations
RR g = 0.0195·M0.55 0.55, (4)(4)
g = 0.0195·M ,
Rg = 0.0195·M 0.55, (4)
where M is the molecular weight of PVP [62];
where M
where M is
is the
the molecular
molecular weight
weight of of PVP
PVP [62];
[62];
Rg = (L·N/6) 0.5, (5)
·N/6) 0.5
RgR=g =(L(L·N/6) 0.5,
, (5)
(5)
where L is the bond length in the polymer chain, nm; N is the number of the bonds [63].
where L is the bond length in the polymer chain, nm; N is the number of the bonds [63].
where L is the bond length in the polymer chain, nm; N is the number of the bonds [63].
50
0
45
50
0
40
45
35 7.0

ζ-potential, mV
40
Intensity, %

30
35 7.0

ζ-potential, mV
5.0
Intensity, %

-5
25
30
20
25
5.0
3.0 -5

15
20 3.0
1.0
10
15
1.0 -10
5
10 0.1
0 0.05 -10
5 0.1
0 1 2 3 4 5 6 7 8
0 1 10 100 1000 0.05
Hydrodynamic diameter, nm 0 1
PVP
2
concentration,
3 4 5
% 6 7 8
1 10 100 1000
Hydrodynamic diameter, nm PVP concentration, %
(a) (b)
(a) (b)
Figure 11. PVP particles size distribution (the curves are marked by PVP concentration in wt.%) (a)
Figure
Figure
and 11. PVP
the PVP particles
particles
particles size
sizedistribution
surface distribution
charge (the
(thecurves
in aqueous curvesare marked
are
solutions marked
(b). byby
PVP concentration
PVP in wt.%)
concentration (a) and
in wt.%) (a)
the PVP
and particles
the PVP surface
particles charge
surface in aqueous
charge solutions
in aqueous (b). (b).
solutions
It is seen that a concentration increase reduces the fraction of expanded polymer chains, and the
degree It is
It is seen
of seen that aa concentration
that
coil swelling concentration
becomes lower, increase
increase
which reduces
reduces the
leadsthe fraction
to fraction
the coil of of expandedand
expanded
contraction polymer
polymer chains,
chains,
decrease and
and
in its the
the
size
degree
(Figure of
degree 11a). coil
of coil swelling
swelling
The PVP coil becomes
becomes lower, which
lower,iswhich
contraction leads to
leads toby
accompanied the coil
theancoil contraction
contraction
increase in theandand decrease
decrease
absolute in its
in its
values size
size
of the
(Figure
(Figure 11a).
11a). The
The PVP
PVP coil
coil contraction
contraction is
is accompanied
accompanied by
by an
an increase
increase in
in
ζ-potential (Figure 11b), which agrees well with the results of quantum-chemical calculations [38], the
the absolute
absolute values
values of
of the
the
ζ-potential
according (Figure
ζ-potentialto(Figure 11b),
which 11b), which agrees
which agrees
the negative well
chargewell with the
with the
is caused results
by results of quantum-chemical
of quantum-chemical
the formation calculations
calculations
of H-bonded complexes [38],
of [38],
the
according to
according
pyrrolidone toring
which
which
with the
the negative
negative
water chargeand
charge
molecules is caused
is caused
shouldby by the formation
the
become formation
lower (more of H-bonded
of H-bonded complexes
positive)complexes of the
of
at dehydration. the
pyrrolidone
pyrrolidone
The PVP adsorptionring with
ring withontowater
water molecules
CNCmolecules and
particles isand should become
should
accompanied become lowerin(more
lower
by growth (more positive) absolute
positive)
the ζ-potential at dehydration.
at dehydration.
value,
The PVP
Thedoes
but PVPnot adsorption
adsorption onto CNC
ontoaffect
significantly CNC theparticles
particles
CNC is is accompanied
accompanied
particle by growth
by growth
size (Figure in the ζ-potential absolute
12a,b).in the ζ-potential absolute value, value,
but does not significantly affect the CNC particle size
but does not significantly affect the CNC particle size (Figure 12a,b).(Figure 12a,b).

0
30

0
30 -10

-10
25:1 -20
20
ζ-potential, mV
% %

25:1
10:1 -20
20 -30
Intensity,

ζ-potential, mV

5:1
10:1
-30
Intensity,

-40
10 5:1
1.5:1
-40
10 1.5:1
0.7:1 -50

0.7:1 -50
0 neat CNC -60

0 neat CNC -600 20 40 60 80 100


1 10 100 1000
0 20
CNC/PVP
40
ratio,
60
% 80 100
1
Hydrodynamic
10
diameter,
100
nm 1000
CNC/PVP ratio, %
Hydrodynamic diameter, nm
(a) (b)
(a) (b)
Figure12.
Figure CNCparticles
12.CNC particlessize
sizedistribution
distributionininthe
thepresence
presenceofofPVPPVP(the
(thecurves
curvesare
aremarked
markedbybythe
the
Figure
PVP/CNC
PVP/CNC 12.concentration
CNC particles
concentration size
ratio)
ratio) distribution
(a) (a)
andand
the theinCNC
CNC the presence
particles
particles of PVP
surface
surface (the
in curves
charge
charge are marked
in aqueous
aqueous by(the
the
suspensions
suspensions
PVP/CNC
(the CNC
CNC concentration
concentration
concentration ratio)
of 0.07of 0.07(a)
wt.%) inand
wt.%) the
thein CNC
the particles
presence
presence of PVP surface
of (b).
PVP (b). charge in aqueous suspensions (the
CNC concentration of 0.07 wt.%) in the presence of PVP (b).
Nanomaterials 2018, 8, 1011 15 of 21

It is usually believed that PVP molecules tend to be adsorbed onto nanoparticles in a linear
open-chain-like pattern as opposed to a compact coiled structure, i.e., the molecular coils are flattened
at the surface [64]. Nevertheless, the surface features of the solid adsorbents can affect the properties of
the adsorbed layer. Hydrogen bonding, electrostatic attraction, or hydrophobic interactions can highly
influence the adsorption behavior [63]. One can hypothesize that strong adsorption due to the high
affinity of PVP for the adsorbent surface causes the PVP molecules to be adsorbed in an open-chain-like
pattern. On the other hand, if the PVP does not interact well with the adsorbent surface, the PVP
molecules tend to be adsorbed in more compact coiled shaped structures.
For example, atomic force microscopy was used by Verma et al. for research the interaction of PVP
with ibuprofen [65]. The obtained AFM images showed that PVP tended to be adsorbed in a compact
coiled structures, demonstrating a lack of affinity for the ibuprofen surface. In addition, the authors
observed patchy preferential adsorption of PVP on one face of the ibuprofen crystal as compared to
the other.
Kotel’nikova et al. [66] studied the effects of PVP molecular weight and aqueous solution
concentration of PVP on adsorption upon microcrystalline cellulose. The authors observed a noticeable
increase in the amount of adsorbed PVP with an increase in its molecular weight, and at higher solution
concentration and temperature. The authors classified the obtained two-step S-shaped isotherm as an
isotherm of type IV and drew a conclusion about the polymolecular character of the PVP adsorption
upon the porous surface of microcrystalline cellulose. Besides, the authors noted that PVP with a
larger molecular weight is both more easily adsorbed and desorbed. Apparently, this indicates that the
polymer is adsorbed in compact globular structures rather than in a flattened conformation as longer
linear chains with a larger molecular mass should bond to the adsorbent surface more firmly.
Hirai and Yakura [62] studied PVP adsorption on palladium (Pd) nanoparticles in methanol.
They determined that the thickness of PVP adsorbed layer on Pd nanoparticles as well as the PVP
radius of gyration increased proportionally to Mw0.55 , where Mw was an averaged molecular weight
of PVP. According to the authors, the adsorbed layer thickness is almost completely determined by the
particular segmental sequence of the adsorbed polymer chain which takes the conformation similar to
that of the free polymer in solution. Consequently, a polymer coiled conformation in solution should
lead to the polymer adsorption in a globule-like structure.
Thus, analyzing the data obtained, one can suppose that PVP is adsorbed onto CNC in a
globule-like structure rather than in a flattened conformation and due to electrostatic attraction
rather than via hydrogen bonding. Nevertheless, the situation can be more complicated because of
PVP strong binding via hydrogen bonding to the (200) facet of the CNC particle which may result in a
flattened conformation of the adsorbed PVP molecules. Computer simulation could reveal and prove
the appropriate mechanism of PVP adsorption onto different facets of CNC particles. However, this
will be the aim of our further research.

3.10. Dispersibility of Freeze-Dried PVP/CNC Composites in Water and Some Organic Solvents
As a PVP molecule contains a strongly hydrophilic pyrrolidone moiety and a considerable
hydrophobic alkyl group, adsorbed PVP can prevent aggregation of nanoparticles through repulsive
forces which arise from the hydrophobic carbon chains. PVP can provide a steric hindrance that will
allow the dried PVP/CNC to be readily redispersed in aqueous systems [27,67,68].
Thus, PVP adsorption onto the CNC surface can lead to a steric hindrance preventing hydrogen
bonding between the CNC particles, which can impart good redispersibility of dried PVP/CNC in
water. To test whether irreversible CNC particle aggregation occurred after drying, we redispersed
freeze-dried PVP/CNC composites in water and some organic solvents. The redispersibility of the
dried PVP/CNC aerogels in water was evaluated in terms of particle-size distribution (Figures 13
and 14) and colloidal stability of the suspensions (Figure S10).
5

Inten

ζ-pote
4 -40
10
3
-50
2
1
-60
0
PVP
0 20 40 60 80 100
1
Nanomaterials
Nanomaterials2018,
2018, FOR10PEER REVIEW
8,8,x1011 100 1000
CNC content in the composite aerogel, %
1616ofof21
21
Hydrodynamic diameter, nm

(a) 0
(b)
30
Figure 13. (a) Particle size distribution of the PVP/CNC aerogels
-10 in water (for comparison, the curves

for the neat PVP and CNC are shown): 1CNC - PVP/CNC-4.6
aero -20
aero; 2 - PVP/CNC-10.9 aero; 3 -
aero; 4 - PVP/CNC-19.6 aero;7 5 - PVP/CNC-28.9 aero; 6 - PVP/CNC-37.9 aero; 7 -

ζ-potential, mV
PVP/CNC-16.3
20
Intensity, %

6 -30
PVP/CNC-54.5 aero. (b) Surface charge of 5PVP coated CNC particles after redispersion of the
PVP/CNC aerogels in water. 4 -40
10
3
-50
2
Figure S10 displays the photographs showing
1
the suspension stability for all the PVP/CNC
-60
samples. The
0 photographs of the suspensionsPVP in the testing vials are taken some time after
0 20 40 60 80 100
ultrasonication. All 1 the CNC
10 suspensions
100 1000 are turbid and no sedimentation can be observed at the
CNC content in the composite aerogel, %
Hydrodynamic diameter, nm
bottom of the vials at the initial stage. As time passes, lots of sedimentation is flocculated in
propanol, chloroform, and (a) dioxane, whereas aqueous suspensions retain their (b) good stability even a
month later.
13.(a) (a)
Figure13.
Figure Particle
Particle size distribution
size distribution of the of the PVP/CNC
PVP/CNC aerogels aerogels in water
in water (for (for comparison,
comparison, the curves
Figures 13a and 14 show particle size distribution of the PVP/CNC aerogels in water and
the the
for curves
neatforPVP
the neat
and PVPCNCand areCNC are shown):
shown): 1—PVP/CNC-4.6
1 - PVP/CNC-4.6 aero; 2aero; 2—PVP/CNC-10.9
- PVP/CNC-10.9 aero; aero;
3 -
propanol, respectively. Figure
3—PVP/CNC-16.3
13b demonstrates surface charge of PVP coated CNC particles after
PVP/CNC-16.3 aero;aero; 4—PVP/CNC-19.6
4 - PVP/CNC-19.6 aero;aero; 5—PVP/CNC-28.9
5 - PVP/CNC-28.9 aero;aero; 6—PVP/CNC-37.9
6 - PVP/CNC-37.9 aero;aero;
7 -
redispersion of the PVP/CNC aerogels in water.
7—PVP/CNC-54.5 aero. (b) Surface charge of PVP coated CNC particles after redispersion of the
PVP/CNC-54.5 aero. (b) Surface charge of PVP coated CNC particles after redispersion of the
PVP/CNCaerogels
PVP/CNC aerogelsininwater.
water.
60

Figure S10 displays the photographs showing the suspension stability for all the PVP/CNC
CNC aero
50
samples. The photographs of the suspensions in the testing vials 7 are taken some time after
40
ultrasonication. All the CNC suspensions are turbid and no sedimentation
6 can be observed at the
Intensity, %

bottom of the vials at the initial


30 stage. As time passes, lots of 5 sedimentation is flocculated in
propanol, chloroform, and dioxane,
20 whereas aqueous suspensions4retain their good stability even a
3
month later. 10 2
1
Figures 13a and 14 show particle size distribution of the PVP/CNC aerogels in water and
0
PVP
propanol, respectively. Figure 13b demonstrates surface charge of PVP coated CNC particles after
1 10 100 1000
redispersion of the PVP/CNC aerogels inHydrodynamic
water. diameter, nm

Figure 14. Particle size distribution of the PVP/CNC aerogels in propanol (for comparison,
Figure 14. Particle size distribution of the PVP/CNC aerogels in propanol (for comparison, the curves
the curves for the neat PVP and 60 CNC are shown): 1—PVP/CNC-4.6 aero; 2—PVP/CNC-10.9 aero;
for the neat PVP and CNC are shown): 1 - PVP/CNC-4.6 aero; 2 - PVP/CNC-10.9 aero; 3 -
3—PVP/CNC-16.3 aero; 4—PVP/CNC-19.6
50 aero; 5—PVP/CNC-28.9CNC aero;
aero6—PVP/CNC-37.9 aero;
PVP/CNC-16.3 aero; 4 - PVP/CNC-19.6 aero; 5 - PVP/CNC-28.9 aero; 76 - PVP/CNC-37.9 aero; 7 -
7—PVP/CNC-54.5 aero.
PVP/CNC-54.5 aero. 40
6
Intensity, %

Figure S10 displays the photographs 30 showing the suspension5 stability for all the PVP/CNC
Figures The
samples. 12 and 13 show of
photographs thatthethe particle size
suspensions in and surface vials
the testing 4charge areafter
takenredispersing
some time the after
20
PVP/CNC aerogels in water remain
ultrasonication. All the CNC suspensions practically unchanged compared
are turbid and no sedimentation 3 to the initial values for
can be observed at thethe
10 2
aqueous
bottom of never-dried
the vials at theCNC in stage.
initial the presence of PVP.
As time passes, lots The redispersed
of sedimentation1 isaqueous
flocculated suspensions
in propanol,
PVP
demonstrate high colloidal stability 0 after 1 month of holding (Figure S10).
chloroform, and dioxane, whereas aqueous suspensions retain their good stability even a month later.
The redispersibility
Figures 13 and 14 show of PVP/CNC
particle sizeaerogels
1 in
10 propanol
distribution
100 is 1000
not as effective.
of the PVP/CNC aerogelsIfinthe PVPand
water content in
propanol,
Hydrodynamic diameter, nm
the composite is less than 60%, aggregates of 1 μm in size are formed
respectively. Figure 13b demonstrates surface charge of PVP coated CNC particles after redispersion (Figure 14). However, the
redispersed
of the suspensions
PVP/CNC aerogelswithin awater.
high PVP content are stable in propanol for 10 days (Figure S10).
Figure 14. Particle size distribution of the PVP/CNC aerogels in propanol (for comparison, the curves
Figures
for 12 and
the neat PVP 13and
show CNCthat are
the shown):
particle size and surface charge
1 - PVP/CNC-4.6 aero; 2after redispersing the
- PVP/CNC-10.9 PVP/CNC
aero; 3 -
aerogels in water aero;
PVP/CNC-16.3 remain practically unchanged
4 - PVP/CNC-19.6 compared toaero;
aero; 5 - PVP/CNC-28.9 the initial values for aero;
6 - PVP/CNC-37.9 the aqueous
7 -
never-dried CNC aero.
PVP/CNC-54.5 in the presence of PVP. The redispersed aqueous suspensions demonstrate high
colloidal stability after 1 month of holding (Figure S10).
Figures 12 and 13 ofshow
The redispersibility PVP/CNCthat the particle
aerogels size andis surface
in propanol charge Ifafter
not as effective. redispersing
the PVP content inthe
the
PVP/CNC aerogels in water remain practically unchanged compared to
composite is less than 60%, aggregates of 1 µm in size are formed (Figure 14). However, the redispersed the initial values for the
aqueous
suspensions never-dried
with a high CNC in the are
PVP content presence
stable inofpropanol
PVP. The for 10 redispersed
days (Figureaqueous S10). suspensions
demonstrate high colloidal stability after 1 month of holding (Figure
In chloroform and dioxane, the redispersed particles of PVP/CNC aerogels form aggregates S10).
The
larger redispersibility
than 3 µm. However, of PVP/CNC aerogels in
these suspensions propanol
are stable for is some
not astime.effective. If the PVP
It is worth content
noting in
that the
the composite
redispersed is less thanbehave
suspensions 60%, aggregates
differently ofin 1chloroform
μm in sizeand aredioxane.
formed If (Figure 14). However,
the former suspension theis
redispersed suspensions with a high PVP content are stable in propanol for 10 days (Figure S10).
Nanomaterials 2018, 8, 1011 17 of 21

most stable at minimum and maximum PVP dosages, the latter, on the contrary, is most unstable under
the same conditions (Figure S10).
These results testify that the enhanced redispersibility of dried CNC/PVP is not due to the
electrostatic mechanism only, rather is specified by both the electrostatic repulsion and the steric
hindrance from the PVP that are adsorbed onto the surface of CNC particles [69].
Figure S11 illustrates PVP particles size distribution in propanol and chloroform (0.2 wt.%
concentration). It is evident that compared to aqueous solutions, the hydrodynamic diameters of PVP
particles in propanol and especially in chloroform tend to increase, probably due to the polymer coils
extension into linear chains. It would be interesting to reveal the relationship of PVP conformations
in various solvents (linear chain, coil, condensed globula) with the possibility to stabilize CNC
suspensions in these solvents. However, this aim is well beyond the scope of the article and will be
achieved in the future work. The specific application of such a system could involve the development
of an effective drug delivery system. The aim of the further research is to enhance physiochemical
properties as well as bioavailability of poorly water-soluble drugs, through preparation of freeze-dried
optimized formulations based on PVP and CNC as hydrophilic carriers.

4. Conclusions
The work describes the preparation of PVP/CNC composite films and aerogels with extensive
characterization. The composites morphology and PVP-assisted CNC self-assembly are explored.
The redispersibility of the aerogels prepared by freeze-drying is investigated in detail. In this work,
we have discovered PVP-assisted CNC self-assembly which produces CNC aggregates with a high
aspect ratio. It was revealed that the particle size of the CNC aggregates after redispersing in water
returned to the initial values for the aqueous never-dried CNC. The redispersed aqueous suspensions
demonstrated high colloidal stability after 1 month of holding. It was shown that the redispersibility
of CNC in an organic solvent (propanol, dioxane, chloroform) could be significantly improved using a
PVP content of 50 wt.%.

Supplementary Materials: The following are available online at http://www.mdpi.com/2079-4991/8/12/1011/


s1, Table S1: CNC characteristics, Figure S1: TEM image of CNC, Figure S2: TG and DTG curves of the PVP/CNC
composite films, Figure S3: DSC traces for heating and cooling of the PVP/CNC composite films, Figure S4:
Typical stress-strain curves of the PVP/CNC composite films, Figure S5: SEM images of the neat PVP aerogel and
the PVP/CNC composite aerogels, Figures S6–S9: SEM images of PVP/CNC composite aerogels for the CNC
particle size estimation, Figure S10: Photos of redispersed suspensions during storage, Figure S11: PVP particles
size distribution in propanol and chloroform.
Author Contributions: Conceptualization, M.V. and O.S.; Methodology, M.V. and O.S.; Validation, M.V., O.S.,
and A.Z.; Formal analysis, M.V., O.S., and A.Z.; Investigation, M.V., N.R., N.K., and A.A.; Data curation, M.V. and
O.S.; Writing—original draft preparation, O.S.; Writing—review and editing, M.V., O.S., and A.Z.; Supervision,
A.Z..; Project administration, A.Z.; Funding acquisition, A.Z. All authors approved the manuscript.
Funding: This research was supported by the Russian Science Foundation, grant number 17-13-01240.
Acknowledgments: The authors would like to thank The Upper Volga Region Centre of Physicochemical Research
(Ivanovo, Russia) and Centre for collective use of scientific equipment of Ivanovo State University of Chemistry
and Technology (Russia) for some measurements carried out using the centers’ equipment.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Zhu, H.; Luo, W.; Ciesielski, P.N.; Fang, Z.; Zhu, J.Y.; Henriksson, G.; Himmel, M.E.; Hu, L. Wood-Derived
Materials for Green Electronics, Biological Devices, and Energy Applications. Chem. Rev. 2016, 116, 9305–9374.
[CrossRef] [PubMed]
2. Moon, R.J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J. Cellulose nanomaterials review: structure,
properties and nanocomposites. Chem. Soc. Rev. 2011, 40, 3941–3994. [CrossRef] [PubMed]
Nanomaterials 2018, 8, 1011 18 of 21

3. Khattab, M.M.; Abdel-Hady, N.A.; Dahman, Y. Cellulose nanocomposites: Opportunities, challenges,


and applications. In Cellulose-Reinforced Nanofibre Composites; Jawaid, M., Boufi, S., Abdul Khalil, H.P.S., Eds.;
Woodhead Publishing: Toronto, ON, Canada, 2017; Chapter 21; pp. 483–516. ISBN 978-0-08-100957-4.
4. Julkapli, N.M.; Bagheri, S. Progress on nanocrystalline cellulose biocomposites. React. Funct. Polym. 2017,
112, 9–21. [CrossRef]
5. Tardy, B.L.; Yokota, S.; Ago, M.; Xiang, W.; Kondo, T.; Bordes, R.; Rojas, O.J. Nanocellulose–surfactant
interactions. Curr. Opin. Colloid In. 2017, 29, 57–67. [CrossRef]
6. Salas, C.; Nypelö, T.; Rodriguez-Abreu, C.; Carrillo, C.; Rojas, O.J. Nanocellulose properties and applications
in colloids and interfaces. Curr. Opin. Colloid In. 2014, 19, 383–396. [CrossRef]
7. Moreau, C.; Villares, A.; Capron, I.; Cathala, B. Tuning supramolecular interactions of cellulose nanocrystals
to design innovative functional materials. Ind. Crop. Prod. 2016, 93, 96–107. [CrossRef]
8. Grishkewich, N.; Mohammed, N.; Tang, J.; Tam, K.C. Recent advances in the application of cellulose
nanocrystals. Curr. Opin. Colloid In. 2017, 29, 32–45. [CrossRef]
9. Tang, J.; Sisler, J.; Grishkewich, N.; Tam, K.C. Functionalization of cellulose nanocrystals for advanced
applications. J. Colloid Interf. Sci. 2017, 494, 397–409. [CrossRef]
10. Niinivaara, E.; Faustini, M.; Tammelin, T.; Kontturi, E. Mimicking the Humidity Response of the Plant Cell
Wall by Using Two-Dimensional Systems: The Critical Role of Amorphous and Crystalline Polysaccharides.
Langmuir 2016, 32, 2032. [CrossRef]
11. Sanchez-Salvador, J.L.; Balea, A.; Monte, M.C.; Blanco, A.; Negro, C. Study of The Reaction Mechanism to
Produce Nanocellulose-Graft-Chitosan Polymer. Nanomaterials 2018, 8, 883. [CrossRef]
12. Tenhunen, T.-M.; Pöhler, T.; Kokko, A.; Orelma, H.; Schenker, M.; Gane, P.; Tammelin, T. Enhancing the
Stability of Aqueous Dispersions and Foams Comprising Cellulose Nanofibrils (CNF) with CaCO3 Particles.
Nanomaterials 2018, 8, 651. [CrossRef] [PubMed]
13. Koczkur, K.M.; Mourdikoudis, S.; Polavarapu, L.; Skrabalak, S.E. Polyvinylpyrrolidone (PVP) in nanoparticle
synthesis. Dalton Trans. 2015, 44, 17883–17905. [CrossRef] [PubMed]
14. Kyrychenko, A.; Korsun, O.M.; Gubin, I.I.; Kovalenko, S.M.; Kalugin, O.N. Atomistic Simulations of Coating
of Silver Nanoparticles with Poly(vinylpyrrolidone) Oligomers: Effect of Oligomer Chain Length. J. Phys.
Chem. C 2015, 119, 7888–7899. [CrossRef]
15. Al-Saidi, W.A.; Feng, H.; Fichthorn, K.A. Adsorption of Polyvinylpyrrolidone on Ag Surfaces: Insight into a
Structure-Directing Agent. Nano Lett. 2012, 12, 997–1001. [CrossRef] [PubMed]
16. Nathanael, A.J.; Seo, Y.H.; Oh, T.H. PVP Assisted Synthesis of Hydroxyapatite Nanorods with Tunable
Aspect Ratio and Bioactivity. J. Nanomater. 2015, 621785. [CrossRef]
17. Kadlubowski, S. Radiation-induced synthesis of nanogels based on poly(N-vinyl-2-pyrrolidone)—A review.
Radiat. Phys. Chem. 2014, 102, 29–39. [CrossRef]
18. An, J.-C.; Weaver, A.; Kim, B.; Barkatt, A.; Poster, D.; Vreeland, W.N.; Silverman, J.; Al-Sheikhly, M.
Radiation-induced synthesis of poly(vinylpyrrolidone) nanogel. Polymer 2011, 52, 5746–5755. [CrossRef]
19. Teodorescu, M.; Morariu, S.; Bercea, M.; Săcărescu, L. Viscoelastic and structural properties of poly(vinyl
alcohol)/poly(vinylpyrrolidone) hydrogels. RSC Adv. 2016, 6, 39718–39727. [CrossRef]
20. Nho, Y.C.; Park, K.R. Preparation and Properties of PVA/PVP Hydrogels Containing Chitosan by Radiation.
J. Appl. Polym. Sci. 2002, 85, 1787–1794. [CrossRef]
21. Yu, H.; Xu, X.; Chen, X.; Hao, J.; Jing, X. Medicated Wound Dressings Based on Poly(vinyl alcohol)/
Poly(N-vinyl pyrrolidone)/Chitosan Hydrogels. J. Appl. Polym. Sci. 2006, 101, 2453–2463. [CrossRef]
22. Halake, K.; Birajdar, M.; Kim, B.S.; Bae, H.; Lee, C.C.; Kim, Y.J.; Kim, S.; Kim, H.J.; Ahn, S.; An, S.Y.; Lee, J.
Recent application developments of water-soluble synthetic polymers. J. Ind. Eng. Chem. 2014, 20, 3913–3918.
[CrossRef]
23. Nešić, A.; Ružić, J.; Gordić, M.; Ostojić, S.; Micić, D.; Onjia, A. Pectin-polyvinylpyrrolidone films:
A sustainable approach to the development of biobased packaging materials. Compos. Part B-Eng. 2017, 110,
56–61. [CrossRef]
24. Balgis, R.; Murata, H.; Goi, Y.; Ogi, T.; Okuyama, K.; Bao, L. Synthesis of Dual-Size Cellulose–
Polyvinylpyrrolidone Nanofiber Composites via One-Step Electrospinning Method for High-Performance
Air Filter. Langmuir 2017, 33, 6127–6134. [CrossRef] [PubMed]
Nanomaterials 2018, 8, 1011 19 of 21

25. Poonguzhali, R.; Basha, S.K.; Kumari, V.S. Synthesis and characterization of chitosan-PVP-nanocellulose
composites for in-vitro wound dressing application. Int. J. Biol. Macromol. 2017, 105, 111–120. [CrossRef]
[PubMed]
26. Wu, X.; Tang, J.; Duan, Y.; Yu, A.; Berry, R.M.; Tam, K.C. Conductive cellulose nanocrystals with high cycling
stability for supercapacitor applications. J. Mater. Chem. A 2014, 2, 19268–19274. [CrossRef]
27. Going, R.J.; Sameoto, D.E.; Ayranci, C. Cellulose Nanocrystals: Dispersion in Co-Solvent Systems and Effects
on Electrospun Polyvinylpyrrolidone Fiber Mats. J. Eng. Fiber. Fabr. 2015, 10, 155–163. [CrossRef]
28. Huang, S.; Zhou, L.; Li, M.-C.; Wu, Q.; Kojima, Y.; Zhou, D. Preparation and Properties of Electrospun
Poly(Vinyl Pyrrolidone)/Cellulose Nanocrystal/Silver Nanoparticle Composite Fibers. Materials 2016, 9,
523. [CrossRef]
29. Hasan, A.; Waibhaw, G.; Tiwari, S.; Dharmalingam, K.; Shukla, I.; Pandey, L.M. Fabrication and
characterization of chitosan, polyvinylpyrrolidone, and cellulose nanowhiskers nanocomposite films for
wound healing drug delivery application. J. Biomed. Mater. Res. A 2017, 105, 2391–2404. [CrossRef]
30. Gao, Y.; Jin, Z. Iridescent Chiral Nematic Cellulose Nanocrystal/Polyvinylpyrrolidone Nanocomposite Films
for Distinguishing Similar Organic Solvents. ACS Sustainable Chem. Eng. 2018, 6, 6192–6202. [CrossRef]
31. Voronova, M.I.; Surov, O.V.; Zakharov, A.G. Nanocrystalline cellulose with various contents of sulfate groups.
Carbohyd. Polym. 2013, 98, 465–469. [CrossRef]
32. Surov, O.V.; Voronova, M.I.; Afineevskii, A.V.; Zakharov, A.G. Polyethylene oxide films reinforced by
cellulose nanocrystals: Microstructure-properties relationship. Carbohyd. Polym. 2018, 181, 489–498. [CrossRef]
[PubMed]
33. Boluk, Y.; Danumah, C. Analysis of cellulose nanocrystal rod lengths by dynamic light scattering and electron
microscopy. J. Nanopart. Res. 2014, 16, 2174. [CrossRef]
34. Fraschini, C.; Chauve, G.; Le Berre, J.-F.; Ellis, S.; Méthot, M.; O’Connor, B.; Bouchard, J. Critical discussion
of light scattering and microscopy techniques for CNC particle sizing. Nord. Pulp Pap. Res. J. 2014, 29, 31–40.
[CrossRef]
35. Ioelovich, M. Characterization of various kinds of nanocellulose. In Handbook of Nanocellulose and Cellulose
Nanocomposites, 1st ed.; Kargarzadeh, H., Ahmad, I., Thomas, S., Dufresne, A., Eds.; Wiley: New York, NY,
USA, 2017; Chapter 2; pp. 51–100. ISBN 978-3-527-33866-5.
36. Zhou, C.; Chu, R.; Wu, R.; Wu, Q. Electrospun polyethylene Oxide/Cellulose nanocrystal composite
nanofibrous mats with homogeneous and heterogeneous microstructures. Biomacromolecules 2011, 12,
2617–2625. [CrossRef] [PubMed]
37. Peresin, M.S.; Habibi, Y.; Zoppe, J.O.; Pawlak, J.J.; Rojas, O.J. Nanofiber composites of polyvinyl alcohol and
cellulose nanocrystals: manufacture and characterization. Biomacromolecules 2010, 11, 674–681. [CrossRef]
[PubMed]
38. Lebedeva, T.L.; Fel’dshtein, M.M.; Platé, N.A.; Kuptsov, S.A. Structure of stable H-bonded poly(N-
vinylpyrrolidone)-water complexes. Polym. Sci. A 2000, 42, 989–1005.
39. Lu, Y.; Armentrout, A.A.; Li, J.; Tekinalp, H.L.; Nanda, J.; Ozcan, S. A cellulose nanocrystal-based composite
electrolyte with superior dimensional stability for alkaline fuel cell membranes. J. Mater. Chem. A 2015, 3,
13350–13356. [CrossRef]
40. Niinivaara, E.; Faustini, M.; Tammelin, T.; Kontturi, E. Water Vapor Uptake of Ultrathin Films of Biologically
Derived Nanocrystals: Quantitative Assessment with Quartz Crystal Microbalance and Spectroscopic
Ellipsometry. Langmuir 2015, 31, 12170–12176. [CrossRef]
41. Mathew, A.P.; Gong, G.; Bjorngrim, N.; Wixe, D.; Oksman, K. Moisture Adsorption Behavior and Its Impact
on the Mechanical Properties of Cellulose Whiskers-Based Polyvinylacetate Nanocomposites. Polym. Eng.
Sci. 2011, 51, 2136–2142. [CrossRef]
42. Hakalahti, M.; Faustini, M.; Boissière, C.; Kontturi, E.; Tammelin, T. Interfacial Mechanisms of Water Vapor
Sorption into Cellulose Nanofibril Films as Revealed by Quantitative Models. Biomacromolecules 2017, 18,
2951–2958. [CrossRef]
43. Kontturi, E.; Tammelin, T.; Oesterberg, M. Cellulose—model films and the fundamental approach. Chem. Soc.
Rev. 2006, 35, 1287–1304. [CrossRef] [PubMed]
44. Kontturi, K.S.; Kontturi, E.; Laine, J. Specific water uptake of thin films from nanofibrillar cellulose. J. Mater.
Chem. A 2013, 1, 13655–13663. [CrossRef]
Nanomaterials 2018, 8, 1011 20 of 21

45. Hakalahti, M.; Mautner, A.; Johansson, L.-S.; Hänninen, T.; Setälä, H.; Kontturi, E.; Bismarck, A.;
Tammelin, T. Direct Interfacial Modification of Nanocellulose Films for Thermoresponsive Membrane
Templates. ACS Appl. Mater. Interfaces 2016, 8, 2923–2927. [CrossRef] [PubMed]
46. Kontturi, E.; Laaksonen, P.; Linder, M.B.; Gröschel, A.H.; Rojas, O.J.; Ikkala, O. Advanced Materials through
Assembly of Nanocelluloses. Adv. Mater. 2018, 30, 1703779. [CrossRef] [PubMed]
47. Boles, M.A.; Engel, M.; Talapin, D.V. Self-Assembly of Colloidal Nanocrystals: From Intricate Structures to
Functional Materials. Chem. Rev. 2016, 116, 11220–11289. [CrossRef] [PubMed]
48. Lin, N.; Dufresne, A. Surface chemistry, morphological analysis and properties of cellulose nanocrystals
with gradiented sulfation degrees. Nanoscale 2014, 10, 5384–5393. [CrossRef]
49. Qin, X.; Xia, W.; Sinko, R.; Keten, S. Tuning Glass Transition in Polymer Nanocomposites with Functionalized
Cellulose Nanocrystals through Nanoconfinement. Nano Lett. 2015, 15, 6738–6744. [CrossRef]
50. Lizundia, E.; Fortunati, E.; Dominici, F.; Vilas, J.L.; León, L.M.; Armentano, I.; Torre, L.; Kenny, J.M.
PLLA-grafted cellulose nanocrystals: Role of the CNC content and grafting on the PLA bionanocomposite
film properties. Carbohyd. Polym. 2016, 142, 105–113. [CrossRef]
51. Kargarzadeh, H.; Mariano, M.; Huang, J.; Lin, N.; Ahmad, I.; Dufresne, A.; Thomas, S. Recent developments
on nanocellulose reinforced polymer nanocomposites: A review. Polymer 2017, 132, 368–393. [CrossRef]
52. Gibson, L.J.; Ashby, M.F. Cellular Solids. In Structure and Properties, 2nd ed.; Cambridge University Press:
Cambridge, UK, 1997; ISBN 9781139878326.
53. French, A.D. Idealized powder diffraction patterns for cellulose polymorphs. Cellulose 2014, 21, 885–896.
[CrossRef]
54. Elazzouzi-Hafraoui, S.; Nishiyama, Y.; Putaux, J.-L.; Heux, L.; Dubreuil, F.; Rochas, C. The Shape and Size
Distribution of Crystalline Nanoparticles Prepared by Acid Hydrolysis of Native Cellulose. Biomacromolecules
2008, 9, 57–65. [CrossRef] [PubMed]
55. Nishiyama, Y.; Langan, P.; Chanzy, H. Crystal Structure and Hydrogen-Bonding System in Cellulose Iβ
from Synchrotron X-ray and Neutron Fiber Diffraction. J. Am. Chem. Soc. 2002, 124, 9074–9082. [CrossRef]
[PubMed]
56. Xiong, R.; Grant, A.M.; Ma, R.; Zhang, S.; Tsukruk, V.V. Naturally-derived biopolymer nanocomposites:
Interfacial design, properties and emerging applications. Mat. Sci. Eng. R 2018, 125, 1–41. [CrossRef]
57. Rojas, O.J.; Lokanathan, A.L.; Kontturi, E.; Laine, J.; Bock, H. The unusual interactions between polymer
grafted cellulose nanocrystal aggregates. Soft Matter 2013, 9, 8965–8973. [CrossRef]
58. Capron, I.; Rojas, O.J.; Bordes, R. Behavior of nanocelluloses at interfaces. Curr. Opin. Colloid In. 2017, 29,
83–95. [CrossRef]
59. Kalashnikova, I.; Bizot, H.; Cathala, B.; Capron, I. Modulation of Cellulose Nanocrystals Amphiphilic
Properties to Stabilize Oil/Water Interface. Biomacromolecules 2012, 13, 267–275. [CrossRef]
60. Munier, P.; Gordeyeva, K.; Bergström, L.; Fall, A.B. Directional Freezing of Nanocellulose Dispersions Aligns
the Rod-Like Particles and Produces Low-Density and Robust Particle Networks. Biomacromolecules 2016, 17,
1875–1881. [CrossRef]
61. Jiang, F.; Hsieh, Y.-L. Super water absorbing and shape memory nanocellulose aerogels from TEMPO-oxidized
cellulose nanofibrils via cyclic freezing–thawing. J. Mater. Chem. A 2014, 2, 350–359. [CrossRef]
62. Hirai, H.; Yakura, N. Protecting Polymers in Suspensions of Metal Nanoparticles. Polym. Adv. Technol. 2001,
12, 724–733. [CrossRef]
63. Sperling, L.H. Introduction to Physical Polymer Science, 4th ed.; John Wiley & Sons, Inc.: Bethlehem, PA, USA,
2006; ISBN 978-0-471-70606-9.
64. Robinson, S.; Williams, P.A. Inhibition of Protein Adsorption onto Silica by Polyvinylpyrrolidone. Langmuir
2002, 18, 8743–8748. [CrossRef]
65. Verma, S.; Huey, B.D.; Burgess, D.J. Scanning Probe Microscopy Method for Nanosuspension Stabilizer
Selection. Langmuir 2009, 25, 12481–12487. [CrossRef] [PubMed]
66. Kotel’nikova, N.E.; Panarin, E.F.; Kudina, N.P. Adsorption of polyvinylpyrrolidone from aqueous solutions
on microcrystalline cellulose. Russ. J. Gen. Chem. 1997, 67, 315–319.
67. Cheng, D.; Wen, Y.; Wang, L.; An, X.; Zhu, X.; Ni, Y. Adsorption of polyethylene glycol (PEG) onto cellulose
nano-crystals to improve its dispersity. Carbohyd. Polym. 2015, 123, 157–163. [CrossRef]
Nanomaterials 2018, 8, 1011 21 of 21

68. Mondal, S. Preparation, properties and applications of nanocellulosic materials. Carbohyd. Polym. 2017, 163,
301–316. [CrossRef] [PubMed]
69. Araki, J. Electrostatic or steric?—Preparations and characterizations of well-dispersed systems containing
rod-like nanowhiskers of crystalline polysaccharides. Soft Matter 2013, 9, 4125–4141. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like