You are on page 1of 12

polymers

Article
Improvement of Mechanical, Hydrophobicity and
Thermal Properties of Chinese Fir Wood by
Impregnation of Nano Silica Sol
Enguang Xu 1 , Yanjuan Zhang 2 and Lanying Lin 1, *
1 Research Institute of Wood Industry, Chinese Academy of Forestry, Beijing 100091, China;
enguangxu@foxmail.com
2 Jixian Honors College, Zhejiang Agriculture and Forestry University, Hangzhou 311300, China;
fdlmy@163.com
* Correspondence: linly@caf.ac.cn; Tel.: +86-10-6288-9100

Received: 26 June 2020; Accepted: 20 July 2020; Published: 23 July 2020 

Abstract: In this paper, a wood-SiO2 composite material was prepared via in-situ polymerization
using vacuum/pressure impregnation technology using commercial scale nano silica sol and Chinese
Fir (Cunninghamia lanceolate (Lamb.) Hook.). Scanning electron microscopy (SEM), Fourier transform
infrared spectroscopy (FTIR), X-ray diffraction (XRD), thermogravimetric analysis (TG), and water
contact angle were used to study the changes in the microstructure and physical and mechanical
properties of this composite. The results showed that silica sol can penetrate and distribute into the
wood cell cavities and surface of cell walls and hence combine with the substances of wood materials.
FTIR results indicated that the –OH groups of wood can polycondense in-situ with silica sol to form
Si–O–C covalent bonds, and amorphous SiO2 formed from Si–O–Si bonds between the –OH groups
of silica sol did not change the crystalline structure of wood cell walls. This in-situ formulating
composite significantly improved the compact microstructure, thermal and mechanical properties,
and hydrophobicity of the composites.

Keywords: in-situ composition; nano silica sol; SiO2 -wood composites; wood cell wall;
vacuum/pressure impregnation

1. Introduction
Wood is one of the oldest materials, mainly composed of cellulose, hemicellulose, and lignin,
and is the only truly renewable material. Wood has excellent mechanical properties, unique decorative
properties, and special environmental characteristics, widely used in architecture and interior decoration.
However, some inherent inferior properties, e.g., lower strength and dimension changes over the
exposing environments and easy to lose mechanical strength under attack by insects, have prevented
efficient uses and exploitation of its full potential. To improve the strength and hardness, dimensional
stability, and antiseptic properties, modifications with suitable chemicals have been employed and
proven to be an attractive way [1,2]. In order to improve the physical and mechanical properties and
weather resistance of wood, different scholars use various inorganic or organic modifiers to treat the
wood [2–4].
However, the toxicity of modified products has caught more and more public attention.
The development of nontoxic wood products has become vitally important [5]. Silicon is a kind
of nontoxic and abundant element, which can exist in inorganic or organic compounds in various
forms, and can be used for wood modification [6,7]. Petrified wood is a typical example, which shows
that organic materials can be transformed into organic-inorganic composite materials under certain

Polymers 2020, 12, 1632; doi:10.3390/polym12081632 www.mdpi.com/journal/polymers


Polymers 2020, 12, 1632 2 of 12

conditions [8]. This type of the SiO2 -wood composite is one of the earliest wood modification methods.
The double diffusion method is used to introduce the modifier into the wood to improve the wood
performance, especially the wood’s fire resistance and decay resistances have been significantly
improved [9–12]. Water glass (sodium silicate) is a common modifier that reacts with metalic salts
to form inorganic deposits. If the sodium ions in water glass are replaced with aluminum sulfate
(Al2 (SO4 )3 ) and calcium chloride (CaCl2 ), the silicate component can be added to the interior of the
wood by impregnation [9]. Other attempts for further improvement included barium chloride (BaCl2 ),
boric acid (H3 BO3 ), borax (Na2 B4 O7 ), boron trioxide (B2 O3 ), and potassium borate (K2 B4 O7 ) [10,12].
These modified woods have high fungal resistance and fire resistance, but poor water resistance,
because various salt components in water glass and the unreacted salt in the lumina of the wood cells
have strong moisture absorption, and these modifiers have poor resistance to bleeding [10].
In order to improve the bleed resistance of inorganic salts, some scholars have begun to use
the sol-gel method to prepare inorganic wood composite [6,7]. The main raw materials used in the
sol-gel process are metal alkoxides, which are used as precursors, such as tetraethoxysilane (TEOS),
2-heptadecafluorooctylethyltrimet-hoxysilane (HFOETMOS), methyltrimethoxysilane (MTMOS),
and polydimethylsiloxane (PDMS) [13–19]. Other reagents have also been investigated as precursors,
including silicon alkoxides, silicon alkoxides derivatives, titanium alkoxides, and titanium alkoxides
derivatives [20–28]. The precursor is dissolved in water or an organic solvent (methanol, ethanol,
isopropanol, 2-methoxyethanol), and an acid or alkali is added as a hydrolysis catalyst so that the
precursor can undergo hydrolysis reaction with the solvent or water in the wood; these hydrolysates
are aggregated into nano-scale particles to produce a sol. The sol is introduced into the wood cell
cavity, cell wall, and pit cavity by coating or vacuum pressure impregnation, then hydrolyzed and
condensed at the same time. After aging, the sol becomes a gel with a three-dimensional network
structure, and these gels are dried to form inorganic nano-oxides, which are deposited in the cell wall
and cell cavity of the wood by hydrogen bonding or physical filling. The obtained modified wood
has improved dimensional stability, fire resistance, and anti-leachability while retaining the porous
structures characteristic of wood.
Nanosols (nano particulate inorganic sols) are another kind of wood modifier. It is a nano-scale
transparent dispersion made by mixing inorganic particles with water or organic liquid. The diameter
of the particles in the dispersion are less than 50 nm and the solid content is between 4% and 20% by
weight [29]. Like other nanomaterials, it has a high surface-to-volume ratio, making it show greater
activity in surface-related phenomena than bulky systems of the same mass. Common treatment
methods include sol-gel, sonochemistry, solvothermal, nanopolymer carriers, combustion, chemical
vapor deposition, and coating treatment. Among them, the sol-gel method is the most widely used
and can introduce metal nanoparticles (gold, copper, and silver) and metal oxides (zinc, aluminum,
and titania) into the wood [29]. However, the sol-gel method is difficult to popularize in industrial
production, because the modifiers used are expensive, the moisture content of wood is strictly controlled,
and the processing technology is complex.
Nano-silica sol is also one of these nanomaterials. There are many different ways to prepare
nano-silica sol. The most commonly used methods are the ion exchange method, one-step hydrolysis
of silica powder, and hydrolysis of silane. The ion exchange process is the most mature and widely
used process. Nano-silica sol can be obtained by removing alkali in water glass using ion exchange
technology. If the reaction continues, the nano-silica sol will eventually polymerize to form particles of
amorphous silicon dioxide. Re-addition of alkali at an appropriate time stops the polycondensation
reaction, so the unmodified nano-silica sol is usually alkaline. However, the nano aqueous silica sols
used in the industry have undergone different surface modifications to become acidic or neutral [29,30].
Compared with other nanomaterials, these products are inexpensive and widely used in the fields
of papermaking, ceramics, catalysts, abrasives, and solid-state electrochemical devices. In addition,
silica sols are used in many coating applications to improve mechanical properties and anti-blocking,
adhesion and wetting properties [31–33]. In addition, the high surface-to-weight ratio (≥200 m2 /g) and
Polymers 2020, 12, 1632 3 of 12

nanoscale particle size of nano sols could be most compatible to the microstructural network system
of wood; they can easily, effectively, and deeply penetrate into the wood substrates and a uniform
distribution can be achieved.
In this study, a vacuum pressure impregnation method was used to introduce a commercial
aqueous nano-silica sol into Chinese fir (Cunninghamia lanceolate (Lamb.) Hook). The applied nano-
silica sol is uniformly filled in wood then nano particles are in situ synthesized on a wood surface by
chemical reactions, to present a wood-SiO2 composite. The structure of the composite materials was
observed, and the thermal and mechanical properties as well as the hydrophobicity of the composites
were evaluated.

2. Materials and Methods

2.1. Materials
Fast-growing Chinese Fir (Cunninghamia lanceolate (Lamb.) Hook.) were collected from a plantation
in Hangzhou, Zhejiang Province, China. Samples were taken from 0.3 m above the ground and then
dried to 30% in a drying kiln. The dried timber was then cut into 300 × 70 × 70 mm blocks according to
GB/T1928~1929-2009; all the samples were taken from the same tree to ensure the minimum difference
in density and mechanical properties. Flat-sawn board was selected for the experiment and then
conditioned in a climate chamber at 23 ± 2 ◦ C and 65 ± 2% relative humidity (RH) until its equilibrium
before silica sol impregnation.
Nano silica sol, an odorless, non-toxic colloid formed using SiO2 nanoparticles in water, whose
molecular formula is mSiO2 .nH2 O, was purchased from Qingdao Hengshengda Chemical Co., Ltd.,
Qingdao, Shandong Province, China. Its pH value was 6.97, the solid content was 21.24%, the specific
gravity was 1.153 g/cm3 , the average size was 20.54 nm, and the dynamic viscosity was 1.87 × 10−3 Pa.s.

2.2. Preparation of SiO2 -Wood Composites


According to the full cell impregnation process (Bethel process), a vacuum pressure impregnation
tank is used to impregnate the wood. After the wood is placed in the cavity of the impregnation tank,
a weight is used to press the wood to avoid the movement of the wood during the impregnation process.
First, the cavity was evacuated and kept for 10 min, with a vacuum of −0.09 MPa. Then, the silica
sol was added into the chamber and pressure impregnation was started. The pressure was 1.0 MPa
and the holding time was 30, 60 and 90 min according to the experimental settings. After reaching the
time, the silica sol was withdrawn from the cavity. Then, negative pressure (−0.09 MPa) was applied
to the cavity for 10 min to remove unnecessary silica sol on the sample surface. After impregnation,
all samples were dried at 103 ◦ C for 24 h and then conditioned under 23 ± 2 ◦ C and 65% relative
humidity. The weight and dimension changes of the wood samples were recorded at each step of
the experiment.

2.3. Characterization of SiO2 -Wood Composites

2.3.1. Weight Percent Gain (WPG)


The mass change of each sample was calculated according to the following formulas with all
measurements taken at oven-dry state:

WPG (%) = (Wt − W0 )/W0 × 100% (1)

where Wt is the mass of sample after treatment and W0 is the mass of sample before treatment.
Polymers 2020, 12, 1632 4 of 12

2.3.2. Scanning Electron Microscopy (SEM) Analysis


Small blocks of 5 (R-radial direction) × 5 (T-tangential direction) × 2 (L-longitudinal direction)
mm in size were cut from the center of untreated and treated samples, and their cross sections were
polished with a microtome. After ion-sputtering with gold, the gold coated portion was studied with
SEM (FEI-XL30, Hillsboro, OR, USA) at 20 keV.

2.3.3. Fourier Transform Infrared Spectroscopy (FTIR) Analysis


Chemical changes and bonding force between the silica and wood were analyzed using FTIR
(Impact 410, Madison, WI, USA). Untreated and treated samples were performed with wave numbers
ranging from 4000 to 400 cm−1 using 50 scans with a resolution of 4 cm−1 .

2.3.4. X-ray Diffraction (XRD) Analysis


The crystal structures of the untreated and treated samples were characterized using XRD (TTR3,
Tokyo, Japan) patterns (Cu Kα radiation, λ = 0.154 nm) operating at 50 kV and 200 mA. The 2θ
ranged from 5◦ to 60◦ with 0.02◦ scanning resolution. The crystallinity index (CrI, %) was determined
as follows:
CrI = (I200 − Iam )/I200 × 100%

where I200 is the adsorption peak intensity of the (200) reflection (at 2θ = 22.3◦ ), and Iam is the minimum
adsorption intensity between (110) and (200) peaks (at 2θ = 18◦ ).

2.3.5. Thermogravimetric (TG) Analysis


Shimadzu DTG-60 (Japan) was used for TG experiments to evaluate its thermal performance.
About 5 mg of the sample was introduced into the sample pan and heated from room temperature to
600 ◦ C with static air as a purge gas at a heating rate of 5 ◦ C/min.

2.3.6. Water Contact Angle Analysis


To evaluate the water-repellent or hydrophobicity properties of the composites, the tangential
and radial contact angles (CAs) of the samples before and after treatment were measured with a CA-W
type automatic contact angle meter. For the measurement of the static contact angle, a droplet volume
of 5 µL was selected. The dynamic contact angle was continuously measured over 20 s.

2.3.7. Mechanical Test


According to the three-point bending test method described in the Chinese national standard
GB/T 1936.1-2009 (method of testing in bending strength of wood), the bending strength and elastic
modulus of the wood before and after treatment were measured. The sample size was 300 (L) × 20 (T)
× 20 (R) mm, and the load loading speed was 3 mm/min. The surface hardness change before and
after the treatment of wood was measured in accordance with the Brinell hardness method in GB/T
1941–2009 (method of testing in hardness of wood). The sample size was 70 (L) × 70 (T) × 50 (R) mm.
A steel ball with a diameter of 5.64 mm was pressed into the tangential section of the wood, and the
loading rate was 5 mm/min until the indentation depth was 2.82 mm. The number of specimens for
each test was selected according to the method in GB/T 1929–2009.

3. Results

3.1. Mass and Microstructure of the Developed Composites


The average change in mass (averages of 24 replicates) of the developed SiO2 -wood composites
at different impregnating times is shown in Figure 1. It can be seen that the mass increased linearly
with the increase of impregnating time (r2 = 0.99). The mean values of SiO2 -wood composites mass
Polymers 2020, 12, 1632 5 of 12

increased from 16.5% to 30.3% (after 90 min impregnation) after curing, with a standard deviation of
3.1% and
Polymers 6.6%
2020, 12, x and with REVIEW
FOR PEER a coefficient of variation of 0.188 and 0.218, respectively. 5 of 11

Polymers 2020, 12, x FOR PEER REVIEW 5 of 11

Weightpercent
Figure1.1.Weight
Figure percentgain
gain(WPG)
(WPG)values
valuesofofsamples
samplestreated
treatedatatdifferent
differentimpregnating
impregnatingtimes
timesunder
under
vacuum/pressure.
vacuum/pressure.
Figure 1. Weight percent gain (WPG) values of samples treated at different impregnating times under
vacuum/pressure.
ItItisisevident
evidentthat thatthe
theuntreated
untreatedsamplesampleisishighly
highlyporous
porous(Figure
(Figure2a). 2a).However,
However,after afteraashort
shorttime
time
(30 min) of silica
(30 min) ofIt silica sol impregnation,
sol impregnation,
is evident that the untreated the cell
thesample walls
cell walls and a few
and aporous
is highly tracheids
few tracheids
(Figure 2a). were
were combined/filled
combined/filled
However, with
after a shortwith silica
timesilica
gel, and
(30 the
min) internal
of silica surfaces
sol of
impregnation, lumensthe were
cell covered
walls and a with
few an
tracheidsuneven
were
gel, and the internal surfaces of lumens were covered with an uneven layer of silica gel. Meanwhile, layer of silica
combined/filled gel.
withMeanwhile,
silica
similar
similargel,toto
andthe
the thenatural porous
internalporous
natural structure
surfacesstructure
of lumensof of
wereuntreated wood,
covered wood,
untreated with anthethe SiO2layer
uneven
SiO -wood
2-wood composite
of silica with30
gel. Meanwhile,
composite with 30 min
min
similar
impregnationalmost
impregnation to the natural
almostmaintainedporous structure
maintainedcellular of untreated
cellularstructure wood,
structure (Figure the
(Figure 2b). SiO
2b). When -wood
Whenthe
2 composite
theimpregnatingwith
impregnatingtime 30 min
timewas was
impregnation
increased, the amount almost
of maintained
SiO gel cellular on
deposited structure
the cell(Figure
wall 2b). Whenand
increased, thethe
impregnating
sample time increased.
weight was
increased, the amount of SiO2 gel deposited on the cell wall increased, and the sample weight
2
increased,
The increase the amount of SiO2 gel deposited on the cell wall increased, and filled
the sample weight
increased. Theinincrease
weight is
inmainly
weightcaused
is mainlyby the increase
caused by the in increase
the content
in theof silica
content in thefilled
of silica cell lumens,
in the
increased. The increase in weight is mainly caused by the increase in the content of silica filled in the
andlumens,
cell the content and of silica
the contentadhered
of to adhered
silica the cell wall
to is cell
the less.wall
Theismicrostructures
less. The of SiO2 -wood
microstructures of composite
SiO 2-wood
cell lumens, and the content of silica adhered to the cell wall is less. The microstructures of SiO2-wood
with 90 minwith
composite impregnating
90 90 minmin is shown in Figure
impregnating 2c; it is apparent thatitsignificant polycondensed SiO2
composite with impregnatingisis shown shown in Figure 2c;
in Figure 2c;it is isapparent
apparent thatthat significant
significant
gels can
polycondensed be
polycondensedobserved
SiO2SiO in
gels the
cancan
2 gels
cell
be belumens.
observed This
observedininthe may
the cell be due
cell lumens.
lumens. This to the
Thismay substantial
maybebe due
due Si–O–Si
to the
to the cross-linking
substantial
substantial Si– Si–
framework
O–Si formed.
cross-linking framework
O–Si cross-linking framework formed. formed.

Figure 2. Cont.
Polymers 2020,
Polymers 12, x1632
2020, 12, FOR PEER REVIEW 66 of
of 11
12

Polymers 2020, 12, x FOR PEER REVIEW 6 of 11

Figure 2.
Figure SEMmicrographs
2. SEM micrographs of
of (a)
(a)untreated
untreated wood
wood and
and the
the SiO
SiO22-wood
-woodcomposites
composites with
with (b)
(b) 30
30 min
min and
and
(c) 90
(c) 90 min
min impregnating
impregnating time.
time.
Figure 2. SEM micrographs of (a) untreated wood and the SiO2-wood composites with (b) 30 min and
3.2. Characteristics and Schematic
(c) 90 min impregnating time. of Chemical Bonding
3.2. Characteristics and Schematic of Chemical Bonding
TheCharacteristics
3.2. FTIR spectra ofSchematic
the treated and untreated samples are shown in Figure 3. It is interesting
The FTIR spectraand of the treated
−1 stretching
of Chemical Bonding samples
and untreated are shown in Figure 3. It is interesting
that the band at 3421 cm−1 vibration of –OH in the untreated wood is replaced by the
that the band
The FTIRat 3421 cmof the
spectra stretching
treated vibration
and of
untreated –OH
samplesin the
are untreated
shown wood
in Figure is isreplaced
3. It interesting by the
stretching vibration band at 3432 cm−1 for SiO2 -wood composite. The band intensity is also decreased.
stretching
that thevibration
band at band
3421 cmat 3432 cm−1 forvibration
−1 stretching
SiO2-wood composite.
of –OH The bandwood
in the untreated intensity is also by
is replaced decreased.
the
These observations indicate
band that thecm –OH reaction groups of wood Themay combine with silica sol to form
Thesestretching vibration
observations indicate atthat
3432 the –OH−1 for SiO 2-woodgroups
reaction
−1
composite.
of wood bandmayintensity
combine is also
−1 (–OH with decreased.
silica sol to
Si–O–C covalent bonds. Bands at 2927 cm (–C–H stretching) and 1596 cm bending) to were
form These
Si–O–C observations
covalent indicate
bonds. Bands that theat–OH 2927reaction groupsstretching)
cm−1 (–C–H of wood may andcombine
1596 cm with silica sol
−1 (–OH bending)
almost unchanged for wood andBandsSiO2 -wood 2927 composite. However, and the Si–O–Si
1596 cm−1bands
(–OH of SiO2 -wood
were form
almostSi–O–C covalent
unchanged for bonds.
wood and at SiO cm−1 (–C–H stretching)
2-wood composite. However, the Si–O–Si bands of SiO2-
bending)
composite gave unchanged
were almost three characteristic
for wood and bands SiO(Figure 3b): a rocking
2-wood composite. However, vibration at 474bands
the Si–O–Si cm−1−1 , aSiO
of bending
2-
wood composite gave −1 three characteristic bands (Figure 3b): a rocking vibration −1 at 474−1cm , a bending
wood composite gave three characteristic bands (Figure 3b): a rocking vibration
vibration at 806 cm−1 , and an asymmetrical stretching vibration at 1110 cm −1 . Due to the influence of at 474 cm , a bending
vibration at 806 cm , and anan asymmetrical stretching vibration
−1 , theat at1110
1110 cm . Duethe to influence
the influence of
broadvibration at 806 cm
and intensive −1, and
Si–O–Si bonds asymmetrical
of silica sol stretching vibration
at 1110 cm –OH cm−1. Due toband
association of and
in cellulose
broadbroad
and andintensive Si–O–Si
intensive Si–O–Si bonds
bondsofofsilica
silica sol at 1110 cm
cm−1, ,the
−1 the –OH association band in cellulose
hemicellulose at 1110 cm −1 overlapped withsol theatSi–O–C
1110 bonds,–OH association
whereby band
the –OH in cellulose
groups of wood
and hemicellulose
and hemicellulose at 1110 cmcm
at 1110
−1 overlapped
−1 overlappedwith with the Si–O–Cbonds,
the Si–O–C bonds, whereby
whereby thethe
–OH –OHgroupsgroups of wood
of wood
may have combined with silica sol to form the Si–O–Si and/or Si–O–C covalent bonds.
may have combined
may have combined withwithsilica
silica solsoltotoform
formthethe Si–O–Si and/orSi–O–C
Si–O–Si and/or Si–O–C covalent
covalent bonds.
bonds.

FigureFigure 3. FTIR
3. FTIR spectra
spectra of of
(a)(a) untreatedwood
untreated wood and the
theSiO
SiO2-wood
2 -woodcomposites withwith
composites (b) 16.5%, (c) 22.8%
(b) 16.5%, (c) 22.8%
and (d) 30.3%
and (d) 30.3% WPGs. WPGs.
Figure 3. FTIR spectra of (a) untreated wood and the SiO2-wood composites with (b) 16.5%, (c) 22.8%
and Further
(d)
Further 30.3% toWPGs.
to the the
FTIRFTIR results,
results, thethe schematicofofSiO
schematic SiO2-wood
-wood composite
compositeformation
formation is shown
is shown in Figure
in Figure 4.
2
4. Under external vacuum and pressure, the impregnating silica sol could flow through wood cells
Under external vacuum and pressure, the impregnating silica sol could flow through wood cells and
and coattothe cell wall to some extent. Vacuum and pressure have important functions in these
coat Further
the cell wall thetoFTIR
someresults,
extent.the schematic
Vacuum andof SiO 2-wood
pressure havecomposite
importantformation
functionsisinshown in Figure
these processes;
4. processes;
Under externalthe vacuum
initial vacuum
and could firstly
pressure, the impregnate
impregnating the wood
silica cell
sol wall quickly
could flow and uniformly.
through wood cells
the initial
Pressure
vacuum
then
could firstly
increases the
impregnate
wood
the wood
impregnation
cell wall
efficiency of
quickly
silica sol
andpromotes
and
uniformly. the
Pressure
chemical
then
and coat the
increases the cell wall
wood to some extent.
impregnation Vacuum
efficiency of and
silica sol pressure
and have the
promotes important
chemical functions
reaction in these
between
reaction between the –OH groups of the silica sol and the –OH groups of cell wall components (Figure
processes; the initial
the –OH groups vacuum
of the could
silica sol andfirstly
the –OHimpregnate
groups ofthecellwood cell wall quickly
wall components and4b)
(Figure uniformly.
and also
Pressure then increases the wood impregnation efficiency of silica sol and promotes the chemical
between the –OH groups of silica sol (Figure 4a). The main reaction sites are the –OH groups of the cell
reaction between the –OH groups of the silica sol and the –OH groups of cell wall components (Figure
Polymers 2020, 12, x FOR PEER REVIEW 7 of 11

4b) and also between the –OH groups of silica sol (Figure 4a). The main reaction sites are the –OH
Polymers
groups2020, 12, 1632
of the cell wall polymer, which includes cellulose, hemicellulose, and lignin. The7 chemical 7 of 12
Polymers 2020, 12, x FOR PEER REVIEW of 11
reaction could also take place in the lumen of the wood. It is worth noting that the aperture of the
lumen 4b)isand
considerably
also between larger thangroups
the –OH the size of thesol
of silica cell wall 4a).
(Figure pores.
TheThemaindeposited silica
reaction sites
wall polymer, which includes cellulose, hemicellulose, and lignin. The chemical reaction could also
aresol
theis–OH
unstable
in thegroups of the cell
cell lumens, andwall
thepolymer, which could
final vacuum includes cellulose,
eliminate thehemicellulose, and lignin.
extra uncondensed Thesol
silica chemical
in lumens.
take place in the lumen of the wood. It is worth noting that the aperture of the lumen is considerably
reaction
Further ovencould
drying also takethe
after place in the process
vacuum lumen offor thethe
wood.
finalItcuring
is worth noting the
ensures thatestablishment
the aperture ofof thethe Si-
largerlumen
than theis size of the cell
considerably wall
larger pores.
than the Theofdeposited
size the cell silica
wall sol is
pores. Theunstable in the
deposited cell
silica sollumens,
is unstableand the
O-wood bonds and Si-O-Si cross-linking framework.
final vacuum
in the cellcould
lumens, eliminate
and the the
finalextra uncondensed
vacuum silicathe
could eliminate solextra
in lumens. Further
uncondensed oven
silica soldrying after the
in lumens.
vacuum process for the final curing ensures the establishment of the Si-O-wood
Further oven drying after the vacuum process for the final curing ensures the establishment of the Si- bonds and Si-O-Si
cross-linking
O-wood bondsframework.
and Si-O-Si cross-linking framework.

Figure 4. Schematic diagram of SiO2-wood composite formation, (a) reaction with –OH groups of
Figure 4. Schematic diagram of SiO -wood composite formation,
formation,(a)(a)reaction with –OH groups of silica
silicaFigure
sol, (b)4.reaction
Schematic diagram
with of 2SiO
the –OH 2-wood
groups ofcomposite
cell wall components . reaction with –OH groups of
silica
sol, (b) sol, (b)with
reaction reaction
the with
–OHthe –OH groups
groups of cellcomponents.
of cell wall wall components .

The chemical
The chemical
bonding
bonding
may
may
result in internal
resultinininternal
restructure
internal restructure
of the
restructure ofofthe
composites. The X-ray diffraction
The chemical bonding may result thecomposites.
composites. TheTheX-ray diffraction
X-ray diffraction
curves of untreated
curves of untreatedwoodwood andandSiO 2-wood composites with different WPGs indicate that the 2θ
SiO 2-wood composites with different WPGs indicate that the 2θ
curves of untreated wood and SiO 2 -wood composites with different WPGs indicate that the 2θ
diffraction curves
diffraction of all
curves of thethe
all samples
samples are
are similar,
similar, and
and theintensity
the intensityisis not
not much
much different
different (Figure
(Figure 5). 5). The
The
diffraction curves of all the samples are similar, and the intensity is not much different (Figure 5).
XRD XRDpatterns for all
patterns for materials
all materialsshow
showthree
threepeaks
peaks at 2θ
2θ values
valuesofof16.7°,
16.7°, 22.8°,
22.8°, andand 34.5°,
34.5°, corresponding
corresponding
The XRD patterns for all materials show three peaks at 2θ values of 16.7◦ , 22.8◦ , and 34.5◦ , corresponding
to theto(110), (200),
the (110), andand
(200), (004) diffraction
(004) diffractionplane
planeofof cellulose Iβpattern
cellulose Iβ pattern [34].
[34]. It can
It can be seen
be seen that that the nano-
the nano-
to the (110), (200), and (004) diffraction plane of cellulose Iβ pattern [34]. It can be seen that the
silicasilica sol impregnation
sol impregnation treatmentdid
treatment didnotnotdestroy
destroy oror even
evenchange
changethethe crystal structure
crystal of the
structure of wood
the wood
nano-silica
samples.sol impregnation treatment diddiffraction
not destroy orintensity
even changerelative
the crystal structure of the wood
samples. WithWith
the the increase
increase of of WPG,the
WPG, the diffraction peak
peak intensityand and relative crystallinity decreased,
crystallinity decreased,
samples. With the increase of WPG, the diffraction peak intensity and relative crystallinity
which may be attributed to the presence of amorphous SiO2 in the cell wall of wood. The crystallinity decreased,
which may be attributed to the presence of amorphous SiO2 in the cell wall of wood. The crystallinity
whichofmay be attributed
untreated wood andtotreated
the presence
wood with of amorphous
16.5%, 22.8%,SiO 2 in
and the weight
30.3% cell wall ofrates
gain wood. wereThe crystallinity
45%, 40%,
of untreated wood and treated wood with 16.5%, 22.8%, and 30.3% weight gain rates were 45%, 40%,
of untreated
37%, andwood and treated wood with 16.5%, 22.8%, and 30.3% weight gain rates were 45%, 40%,
32%, respectively.
37%, and 32%, respectively.
37%, and 32%, respectively.
(200)
(200)
(110)

(110)
(004)
Intensity(a.u)

(004) (a)
Intensity(a.u)

(b)
(a)

(c(
)b)

(d)
(c)

10 20 30 40 50 60
(d)
2θ(°)

Figure 5. XRD patterns of (a)10untreated20wood and


30 the SiO240 50
-wood composite 60 (b) 16.5%, (c) 22.8%
with
2θ(°)
and (d) 30.3% WPGs.
Figure 5. XRD patterns of (a) untreated wood and the SiO22 -wood composite with (b) 16.5%, (c) 22.8%
Figure 5.
and
and (d)
(d) 30.3%
30.3% WPGs.
WPGs.
evaporation of absorbed moisture and the partial decomposition of hemicellulose. An abrupt mass
loss (region 2) occurred at 197–336 °C, which was mainly caused by hemicellulose decomposition,
accompanied the continuous decomposition of cellulose and lignin, and mass loss reached 55.0%.
Starting from 336 °C, all components of the wood gradually degraded leading to aromatization and
carbonization
Polymers 2020, 12, (region
1632 3). After 482 °C, the wood was completely carbonized, and 17.1% carbon 8 of 12
residue was left.
Compared with the untreated wood, the SiO2-wood composites had distinctive TG curves,
3.3. Thermal Properties
characteristically, of SiO
of four 2 -Woodtemperature
different Composites regions: in region 1, a slight mass loss was observed
from TGA
room curves
temperature
are usedto 232 °C, and thethe
to determine mass loss was
thermal almostofthe
stability woodsameandas that
wood ofmaterials.
untreated wood.
As the
The mass lossincreases,
temperature in regionthe 2 decreased rapidly
materials begin from 232and
to pyrolyze °Cthe
to 327 °C, content
residue followed by another
increases, remarkable
which indirectly
mass
means loss (region
that 3) at 327–431
the amount of volatile°C.combustion
The most interesting feature
products may about the
decrease. TGon
Based curves of the
the mass SiO
loss 2-wood
observed
composites was the
in the TG curve, thetemperature
untreated wood from can
431 clearly
°C to 492
be°C (regioninto
divided 4), three
compared withtemperature
different untreated wood,
regionsin
which
(Figurethe 6): mass
from loss
roomdeclined slowly.
temperature The ◦ultimate
to 197 C, only adecomposition
small amount temperature
of mass lossofoccurs
SiO2-wood
in the
composites
material (regionshifted1), towards the higher
and the mass temperature,
loss was about 4.2%,after which
which was lower
mainlyrates of the
due to weight loss were
evaporation of
observed.
absorbed Eventually,
moisture and higher amounts
the partial of residues ofof47.0%,
decomposition 56.5%, andAn
hemicellulose. 69.3% were
abrupt obtained
mass for 16.5,2)
loss (region
22.8 and 30.3
occurred WPG, ◦respectively.
at 197–336 C, which wasThe residue
mainly produced
caused by SiO2 wood
by hemicellulose composites accompanied
decomposition, depends on the the
continuous decomposition of cellulose and lignin, and mass loss reached 55.0%. Starting from 336 ◦ C,
WPG of the composites, and the residual amount increases with the increase of WPG. This result
shows that the silica
all components of thesol impregnated
wood gradually in wood can
degraded withstand
leading higher temperatures
to aromatization and can (region
and carbonization improve 3).
the thermal ◦ properties of wood.
After 482 C, the wood was completely carbonized, and 17.1% carbon residue was left.

1
100 2

197 3
Relative mass loss (%)

232
80 1
4
(d)

60 (c)
327
(b)
431 492
40 336
2

482
20 (a)
3

0 100 200 300 400 500 600


Temperature (℃)
Figure
Figure6.6.Thermogravimetric
Thermogravimetric(TG) (TG)curves
curvesofof(a)
(a)untreated
untreatedwood
woodcontaining
containing(1)(1)slight
slightmass
masslosslossstage,
stage,
(2) abrupt mass loss stage, (3) aromatization and carbonization stage (blue region) and the SiO22-wood
(2) abrupt mass loss stage, (3) aromatization and carbonization stage (blue region) and the SiO -wood
composites
compositeswith
with(b)
(b)16.5%,
16.5%,(c)(c)22.8%, and
22.8%, and(d)(d)
30.3% WPGs,
30.3% WPGs,containing (1) (1)
containing slight mass
slight loss loss
mass stage, (2)
stage,
abrupt mass
(2) abrupt lossloss
mass stage, (3) remarkable
stage, (3) remarkablemass
masslossloss
stage, (4) slowly
stage, mass
(4) slowly lossloss
mass stage (red(red
stage region).
region).

Compared with the untreated wood, the SiO2 -wood composites had distinctive TG curves,
characteristically, of four different temperature regions: in region 1, a slight mass loss was observed
from room temperature to 232 ◦ C, and the mass loss was almost the same as that of untreated wood.
The mass loss in region 2 decreased rapidly from 232 ◦ C to 327 ◦ C, followed by another remarkable
mass loss (region 3) at 327–431 ◦ C. The most interesting feature about the TG curves of the SiO2 -wood
composites was the temperature from 431 ◦ C to 492 ◦ C (region 4), compared with untreated wood,
in which the mass loss declined slowly. The ultimate decomposition temperature of SiO2 -wood
composites shifted towards the higher temperature, after which lower rates of weight loss were
observed. Eventually, higher amounts of residues of 47.0%, 56.5%, and 69.3% were obtained for 16.5,
22.8 and 30.3 WPG, respectively. The residue produced by SiO2 wood composites depends on the WPG
of the composites, and the residual amount increases with the increase of WPG. This result shows that
Polymers 2020, 12, x FOR PEER REVIEW 9 of 11

3.4. Hydrophobicity of Composites


Figure 7 shows the water contact angles of tangential and radial sections of untreated wood and
Polymers 2020, 12,
composites of 1632
differentWPG over a time period of 20 s. It is evident that the instantaneous contact9 of 12

angles of tangential sections of composites increased with increasing WPG (Figure 7a). The instant
contact
the silicaangle increased byin
sol impregnated about
wood10%canfor the composite
withstand higherwith 17% WPG,and
temperatures 25%can
forimprove
that withthe
23% WPG
thermal
and 30% for that
properties of wood. with 30% WPG. It is very interesting that while the contact angle of untreated wood
consistently reduced by the duration of wetting, from 63.8° to 49.2°, the contact angles of composites
is almost
3.4. independent
Hydrophobicity of the wetting time. Therefore, at the end of 20 min wetting, the composites
of Composites
with 23% WPG increased the contact angles by 61% compared to the matched wood materials.
Figure 7 shows the water contact angles of tangential and radial sections of untreated wood and
The instantaneous contact angle of the composite radial section is slightly higher than that of the
composites of different WPG over a time period of 20 s. It is evident that the instantaneous contact
untreated wood (Figure 7b). It is also interesting that the loading of WPG had little effect on the
angles of tangential sections of composites increased with increasing WPG (Figure 7a). The instant
instantaneous contact angles. However, after 20 s of wetting, the water contact angle of untreated
contact angle increased by about 10% for the composite with 17% WPG, 25% for that with 23% WPG
wood decreased from 73.4° to 54.3°. By contrast, the contact angles of composites almost remained
and 30% for that with 30% WPG. It is very interesting that while the contact angle of untreated wood
unchanged throughout the exposure of water wetting. Therefore, at the end of 20 min, the contact
consistently reduced by the duration of wetting, from 63.8◦ to 49.2◦ , the contact angles of composites is
angle of the composite with 30.3% WPG increased by 42% compared to the matched wood. These
almost independent of the wetting time. Therefore, at the end of 20 min wetting, the composites with
results indicate that silica sol impregnated composites can improve the water repellency of the
23% WPG increased the contact angles by 61% compared to the matched wood materials.
materials and have good hydrophobicity.

85
a b
85

80 80

75 75
Contact angle (°)

Contact angle (°)

70 70

65 65
16.5% WPG
22.8% WPG
60 60
30.3% WPG
16.5% WPG
Untreated
55 55 22.8% WPG
30.3% WPG
50 50 Untreated

45 45
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Time (s) Time (s)

Figure 7. Water contact angles of SiO


SiO22-wood composites on tangential (a) and radial (b) sections.

The instantaneous
3.5. Mechanical Propertiescontact angle of the composite radial section is slightly higher than that of
of the Composites
the untreated wood (Figure 7b). It is also interesting that the loading of WPG had little effect on the
In order contact
instantaneous to evaluate theHowever,
angles. reinforcing effect
after 20 s of nano silica
of wetting, the sol,
watermechanical properties
contact angle were
of untreated
measured. The modulus of
◦ elasticity,
◦ bending strength and hardness of composites
wood decreased from 73.4 to 54.3 . By contrast, the contact angles of composites almost remained and untreated
wood are the
unchanged average of
throughout the62, 35, andof62
exposure replicates,
water wetting.respectively,
Therefore, atwhich areofsummarized
the end in Figure
20 min, the contact angle8.
The mean values for modulus of elasticity of the composites at three WPGs
of the composite with 30.3% WPG increased by 42% compared to the matched wood. These results of 16.5%, 22.8%, and
30.3% were
indicate that9.23,
silica10.08, and 11.12 GPa,
sol impregnated with a standard
composites deviation
can improve of 0.78,
the water 1.16, andof1.47
repellency the GPa and with
materials and
a coefficient of variation
have good hydrophobicity. of 0.085, 0.115, and 0.132, respectively, while the modulus of elasticity of the
untreated wood was 7.30 (±0.64) GPa. This means that the stiffness of the 30.3% WPG composite is
increased
3.5. by 52.3%.
Mechanical The bending
Properties strength of the composite with 30.3% WPG was 80.51 (±8.96) MPa,
of the Composites
which is 28.6% higher than the untreated 62.62 (±4.24) MPa. The hardness of composite materials
In order to evaluate the reinforcing effect of nano silica sol, mechanical properties were measured.
increases with the increase of WPG. When WPGs were 16.5%, 22.8%, and 30.3%, the average hardness
The modulus of elasticity, bending strength and hardness of composites and untreated wood are
of composites was 4.20, 4.75, and 5.46 kN, with a standard deviation of 0.48, 0.62, and 0.88 kN and
the average of 62, 35, and 62 replicates, respectively, which are summarized in Figure 8. The mean
with a coefficient of variation of 0.114, 0.131, and 0.161, respectively, which were increased by 43.7%,
values for modulus of elasticity of the composites at three WPGs of 16.5%, 22.8%, and 30.3% were
62.1%, and 86.3%, respectively, compared with the average hardness of untreated wood. The
9.23, 10.08, and 11.12 GPa, with a standard deviation of 0.78, 1.16, and 1.47 GPa and with a coefficient
mechanical properties of composites are linearly related to WPG. The improvement of the mechanical
of variation of 0.085, 0.115, and 0.132, respectively, while the modulus of elasticity of the untreated
properties is mainly due to the impregnation of the silica sol into the wood to strengthen the wood
wood was 7.30 (±0.64) GPa. This means that the stiffness of the 30.3% WPG composite is increased by
cell walls or fill the voids of the wood structure and their combined effects between wood and silica
52.3%. The bending strength of the composite with 30.3% WPG was 80.51 (±8.96) MPa, which is 28.6%
sol through various physical, chemical, and mechanical interactions/bonding.
higher than the untreated 62.62 (±4.24) MPa. The hardness of composite materials increases with the
increase of WPG. When WPGs were 16.5%, 22.8%, and 30.3%, the average hardness of composites was
4.20, 4.75, and 5.46 kN, with a standard deviation of 0.48, 0.62, and 0.88 kN and with a coefficient of
variation of 0.114, 0.131, and 0.161, respectively, which were increased by 43.7%, 62.1%, and 86.3%,
respectively, compared with the average hardness of untreated wood. The mechanical properties of
Polymers 2020, 12, 1632 10 of 12

composites are linearly related to WPG. The improvement of the mechanical properties is mainly due
to the impregnation of the silica sol into the wood to strengthen the wood cell walls or fill the voids of
the wood structure and their combined effects between wood and silica sol through various physical,
chemical, and mechanical interactions/bonding.
Polymers 2020, 12, x FOR PEER REVIEW 10 of 11

8. The mechanical properties of the untreated


Figure 8. untreated wood
wood and
and the
the SiO
SiO22-wood composites.

4. Conclusions
4. Conclusions
In-situ
In-situ SiO
SiO22-wood
-woodcomposites
compositeshave have successfully
successfullybeen developed
been developed by introducing commercial
by introducing nano
commercial
silica
nano sol into
silica solwood. The in-situ
into wood. combination
The in-situ combinationof wood and silica
of wood sol was
and silica solefficient and effective
was efficient with
and effective
the aid of vacuum impregnation, depending on the duration, with the mass
with the aid of vacuum impregnation, depending on the duration, with the mass and the amounts and the amounts of SiOof2
gel of the composites increasing with the increasing of impregnation time.
SiO2 gel of the composites increasing with the increasing of impregnation time. The composites The composites almost
maintained cellularcellular
almost maintained structure, and the
structure, andSiO 2 gel
the SiOwas significantly combined and extended from the
2 gel was significantly combined and extended from
surface
the surface of the cell lumens, while a small portion of silicasol
of the cell lumens, while a small portion of silica solpenetrated
penetratedinto
intoand
anddeveloped
developed bonds
bonds
with
with thethecell walls.
cell walls.TheThe
processing technologies
processing stimulated
technologies the –OH
stimulated thegroups
–OH of wood, of
groups which combined
wood, which
with silica sol
combined with to silica
form the
sol Si–O–Si
to formand theSi–O–C
Si–O–Si covalent bonds, covalent
and Si–O–C resulting bonds,
in the transformation
resulting in theof
the microstructure of wood to composites and hence significant improvements
transformation of the microstructure of wood to composites and hence significant improvements in in the thermal and
mechanical
the thermalproperties of composites.
and mechanical The hydrophobicity
properties of composites.of Thethe developed composites
hydrophobicity is much
of the lower
developed
than that of raw wood materials.
composites is much lower than that of raw wood materials.
Author Contributions:
Author Contributions: TheAll authors
manuscripthavewasread and agree
completed to thecontributions
through published version of the manuscript.
of all authors. The
Data curation,
E.X. and Y.Z.;
manuscript investigation,
was E.X., Y.Z.
completed through and L.L.; methodology,
contributions of all authors.E.X.,
DataY.Z., and L.L.;
curation, E.X. project administration,
and Y.Z.; L.L.;
investigation, E.X.,
supervision,
Y.Z. and L.L.;L.L.; writing—original
methodology, draft,
E.X., Y.Z., andE.X. and
L.L.; Y.Z.; administration,
project writing—reviewL.L.;andsupervision,
editing, L.L.L.L.;
All authors have read
writing—original
and agreed to the published version of the manuscript.
draft, E.X. and Y.Z.; writing—review and editing, L.L.
Funding: This research was supported by the “Nature Science Foundation of China” (no. 31000272).
Funding: This research was supported by the “Nature Science Foundation of China” (no. 31000272).
Conflicts of Interest: The authors declare no conflict of interest.
Conflicts of Interest: The authors declare no conflicts of interest.
References
References
1. Qin, L.Z.; Lin, L.Y.; Fu, F. Microstructural and micromechanical characterization of modified
1. Qin, L.Z.; Lin, L.Y.;
urea-formaldehyde Fu,penetration
resin F. Microstructural
into wood. and micromechanical
Bioresources characterization
2016, 11, 182–194. of modified urea-
[CrossRef]
2. formaldehyde resin penetration into wood. Bioresources 2016, 11, 182–194.
Lin, L.Y.; Fu, F. The composite wood impregnated with silicon sol solution. Adv. Mater. Res. 2012, 466–467,
2. Lin, L.Y.;[CrossRef]
121–126. Fu, F. The composite wood impregnated with silicon sol solution. Adv. Mater. Res. 2012, 466–467,
3. 121–126. D.; Kutnar, A.; Mantanis, G. Wood modification technologies—A review. iForest Biogeosci. For.
Sandberg,
3. Sandberg,
2017, D.; Kutnar,
10, 895–908. A.; Mantanis, G. Wood modification technologies—A review. iForest-Biogeosciences
[CrossRef]
4. For. 2017, 10, 895–908.
Xie, Y.; Fu, Q.; Wang, Q.; Xiao, Z.; Militz, H. Effects of chemical modification on the mechanical properties of
4. Xie, Y.;Eur.
wood. Fu, Q.; Wang,
J. Wood WoodQ.; Prod.
Xiao, 2013,
Z.; Militz, H. Effects
71, 401–416. of chemical modification on the mechanical properties
[CrossRef]
5. of wood.
Wang, B.;Eur.
Zhang,J. Wood Wood
Y.; Tan, H.;Prod.
Gu, 2013, 71, 401–416.
J. Melamine-urea-formaldehyde resins with low formaldehyde emission
5. Wang, B.; Zhang, Y.; Tan, H.; Gu, J. Melamine-urea-formaldehyde
and resistance to boiling water. Pigment Resin Technol. 2019, 48, 229–236. resins[CrossRef]
with low formaldehyde emission
6. and resistance to boiling water. Pigment Resin Technol. 2019, 48, 229–236.
Mai, C.; Militz, H. Modification of wood with silicon compounds. Inorganic silicon compounds and sol-gel
6. Mai, C.; Militz,
systems: A review. H. Modification of wood
Wood Sci. Technol. with
2004, 37,silicon
339–348.compounds.
[CrossRef]Inorganic silicon compounds and sol-gel
7. systems:
Mai, A review.
C.; Militz, Wood Sci. Technol.
H. Modification 2004,with
of wood 37, 339–348.
silicon compounds. Treatment systems based on organic
7. silicon compounds: A review. Wood Sci. Technol. silicon
Mai, C.; Militz, H. Modification of wood with compounds.
2004, 37, Treatment systems based on organic
453–461. [CrossRef]
silicon compounds: A review. Wood Sci. Technol. 2004, 37, 453–461.
8. Kuczumow, A.; Nowak, J.; Kuziola, R.; Jarzebski, M. Analysis of the composition and minerals diagrams
determination of petrified wood. Microchem. J. 2019, 148, 120–129.
9. Furuno, T.; Uehara, T.; Jodai, S. Combination of wood and silicate (I). Impregnation by water glass and
application of aluminum sulfate and calcium chloride as reactants. Mokuzai Gakkaishi 1991, 37, 462–472.
10. Furuno, T.; Shimada, K.; Uehara, T.; Jodai, S. Combinations of wood and silicate (II). Wood-mineral
Polymers 2020, 12, 1632 11 of 12

8. Kuczumow, A.; Nowak, J.; Kuziola, R.; Jarzebski, M. Analysis of the composition and minerals diagrams
determination of petrified wood. Microchem. J. 2019, 148, 120–129. [CrossRef]
9. Furuno, T.; Uehara, T.; Jodai, S. Combination of wood and silicate (I). Impregnation by water glass and
application of aluminum sulfate and calcium chloride as reactants. Mokuzai Gakkaishi 1991, 37, 462–472.
10. Furuno, T.; Shimada, K.; Uehara, T.; Jodai, S. Combinations of wood and silicate (II). Wood-mineral
composites using water glass and reactance of barium chloride, boric acid, and borax and their properties.
Mokuzai Gakkaishi 1992, 38, 448–457.
11. Furuno, T.; Uehara, T.; Jodai, S. Combination of wood and silicate (III). Some properties of wood-mineral
composites using the water glass-boron compound system. Mokuzai Gakkaishi 1993, 39, 561–570.
12. Yamaguchi, H. Preparation and physical properties of wood fixed with silicic acid compounds.
Mokuzai Gakkaishi 1994, 40, 38–845.
13. Saka, S.; Sasaki, M.; Tanahashi, M. Wood-inorganic composites prepared by the sol-gel process (I).
Wood-inorganic composites with porous structure. Mokuzai Gakkaishi 1992, 38, 1043–1049.
14. Saka, S.; Yakake, Y. Wood-inorganic composites prepared by sol-gel process (III). Chemically-modified
wood-inorganic composites. Mokuzai Gakkaishi 1993, 39, 308–314.
15. Ogiso, K.; Saka, S. Wood-inorganic composites prepared by sol-gels process (II). Effect of ultrasonic treatments
on preparation of wood-inorganic composites. Mokuzai Gakkaishi 1993, 39, 301–307.
16. Ogiso, K.; Saka, S. Wood-inorganic composites prepared by sol-gel process (IV). Effects of chemical bonds
between wood and inorganic substances on property enhancement. Mokuzai Gakkaishi 1994, 40, 1100–1106.
17. Miyafuji, H.; Saka, S.; Yamamoto, A. SiO2 -P2 O5 -B2 O3 wood-inorganic composites prepared by metal alkoxide
oligomers and their fire-resisting properties. Holzforschung 1998, 52, 410–416. [CrossRef]
18. Pries, M.; Mai, C. Fire resistance of wood treated with a cationic silica sol. Eur. J. Wood Wood Prod. 2013, 71,
237–244. [CrossRef]
19. Bak, M.; Molnár, F.; Németh, R. Improvement of dimensional stability of wood by silica nanoparticles.
Wood Mater. Sci. Eng. 2019, 14, 48–58. [CrossRef]
20. Broda, M.; Majka, J.; Olek, W.; Mazela, B. Dimensional stability and hygroscopic properties of waterlogged
archaeological wood treated with alkoxysilanes. Int. Biodeterior. Biodegrad. 2018, 133, 34–41. [CrossRef]
21. Mazela, B.; Kowalczuk, J.; Ratajczak, I.; Szentner, K. Moisture content (MC) and multinuclear magnetic
resonance imaging (MRI) study of water absorption effect on wood treated with aminofunctional silane.
Eur. J. Wood Prod. 2014, 72, 243–248. [CrossRef]
22. Papadopoulos, A.N.; Bikiaris, D.N.; Mitropoulos, A.C.; Kyzas, G.Z. Nanomaterials and Chemical
Modifications for Enhanced Key Wood Properties: A Review. Nanomaterials 2019, 9, 607. [CrossRef]
[PubMed]
23. Qin, C.; Zhang, W.B. Antibacterial property of titanium alkoxide/poplar wood composite prepared by sol-gel
process. Mater. Lett. 2012, 89, 101–103. [CrossRef]
24. Mahr, M.S.; Hübert, T.; Sabel, M.; Schartel, B.; Bahr, H.; Militz, H. Fire retardancy of sol–gel derived titania
wood-inorganic composites. J. Mater. Sci. 2012, 47, 6849–6861. [CrossRef]
25. Mahr, M.S.; Hübert, T.; Schartel, B.; Bahr, H.; Sabel, M.; Militz, H. Fire retardancy effects in single and double
layered sol-gel derived TiO2 and SiO2 -wood composites. J. Sol Gel Sci. Technol. 2012, 64, 452–464. [CrossRef]
26. Mahr, M.S.; Hübert, T.; Stephan, I.; Militz, H.; Shabir Mahr, M.; Hübert, T.; Stephan, I.; Militz, H. Decay
protection of wood against brown-rot fungi by titanium alkoxide impregnations. Int. Biodeterior. Biodegrad.
2013, 77, 56–62. [CrossRef]
27. Unger, B.; Bücker, M.; Reinsch, S.; Hübert, T. Chemical aspects of wood modification by sol-gel-derived silica.
Wood Sci. Technol. 2013, 47, 83–104. [CrossRef]
28. Lu, Y.; Feng, M.; Zhan, H. Preparation of SiO2 -wood composites by an ultrasonic-assisted sol-gel technique.
Cellulose 2014, 21, 4393–4403. [CrossRef]
29. Mahltig, B.; Swaboda, C.; Roessler, A.; Böttcher, H. Functionalising wood by nanosol application.
J. Mater. Chem. 2008, 18, 3180–3192. [CrossRef]
30. Greenwood, P.; Gevert, B. Aqueous silane modified silica sols: Theory and preparation. Pigment Resin Technol.
2011, 40, 275–284. [CrossRef]
31. Papadopoulos, A.N.; Taghiyari, H.R. Innovative Wood Surface Treatments Based on Nanotechnology.
Coatings 2019, 9, 866. [CrossRef]
Polymers 2020, 12, 1632 12 of 12

32. Tu, K.; Kong, L.; Wang, X.; Liu, J. Semitransparent, durable superhydrophobic polydimethylsiloxane/SiO2
nanocomposite coatings on varnished wood. Holzforschung 2016, 70, 1039–1045. [CrossRef]
33. Tsvetkova, I.N.; Krasil’nikova, L.N.; Khoroshavina, Y.V.; Galushko, A.S.; Shilova, O.A. Sol-gel preparation of
protective and decorative coatings on wood. J. Sol-Gel Sci. Technol. 2019, 92, 474–483. [CrossRef]
34. French, A.D. Idealized powder diffraction patterns for cellulose polymorphs. Cellulose 2014, 21, 885–896.
[CrossRef]

© 2020 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