You are on page 1of 12

J Nanopart Res (2010) 12:2415–2426

DOI 10.1007/s11051-009-9802-z

RESEARCH PAPER

Effects of iron oxide nanoparticles on polyvinyl alcohol:


interfacial layer and bulk nanocomposites thin film
Zhanhu Guo • Di Zhang • Suying Wei • Zhe Wang • Amar B. Karki •

Yuehao Li • Paul Bernazzani • David. P. Young • J. A. Gomes •


David L. Cocke • Thomas C. Ho

Received: 9 July 2009 / Accepted: 3 November 2009 / Published online: 19 November 2009
Ó Springer Science+Business Media B.V. 2009

Abstract Iron oxide (a-phase) nanoparticles with surface. With PVA attached to the nanoparticle
coercivity larger than 300 Oe have been fabricated at surface, the nanoparticles are found not only to
a mild temperature by an environmentally benign increase the PVA cross-linking with an increase in
method. The economic sodium chloride has been melting temperature but also to enhance the thermal
found to effectively serve as a solid spacer to disperse stability of the PVA. The nanoparticles are observed
the iron precursor and to prevent the nanoparticles to be uniformly dispersed in the polymer matrix.
from agglomeration. Higher ratios of sodium chloride Scanning electron microscopy (SEM) microstructure
to iron nitrate result in smaller nanoparticles (19 nm also shows an intermediate phase with a strong
for 20:1 and 14 nm for 50:1). The presence of interaction between the nanoparticles and the poly-
polyvinyl alcohol (PVA) limits the particle growth mer matrices, arising from the hydrogen bonding
(15 nm for 20:1 and 13 nm for 50:1) and favors between the PVA and hydroxyl groups on the
nanoparticle dispersion in polymer matrices. Obvious nanoparticle surface. The addition of nanoparticles
physicochemical property changes have been favors the cross-linkage of the bulk PVA matrices,
observed with PVA attached to the nanoparticle resulting in a higher melting temperature, and an
enhanced thermal stability of the polymer matrix.

Z. Guo (&)  D. Zhang  Y. Li  J. A. Gomes 


Keywords Nanoparticles 
D. L. Cocke  T. C. Ho
Integrated Composites Laboratory (ICL), Dan F. Smith Polymer nanocomposites  Thermal properties
Department of Chemical Engineering, Lamar University,
Beaumont, TX 77710, USA
e-mail: zhanhu.guo@lamar.edu

S. Wei  P. Bernazzani Introduction


Department of Chemistry and Physics, Lamar University,
Beaumont, TX 77710, USA Antiferromagnetic a-Fe2O3 (hematite) nanoparticles
(NPs) have attracted much interest due to their wide
Z. Wang
Mechanical and Aerospace Engineering Department, potential applications in many areas, such as pigments
University of California Los Angeles, Los Angeles, (Feldmann 2001), catalysts (Glisenti 1998; Ren et al.
CA 90095, USA 2009; Rofer-Depoorter 1981; Xie et al. 2009; Zhang
et al. 2009; Zheng et al. 2007), photocatalysts
A. B. Karki  David. P. Young
Department of Physics and Astronomy, Louisiana State (Kay et al. 2006), gas sensors (Fukazawa et al.
University, Baton Rouge, LA 70803, USA 1993; Huo et al. 2005; Neri et al. 2002; Wang et al.

123
2416 J Nanopart Res (2010) 12:2415–2426

2008), nonlinear optics (Yu et al. 2000), reinforcing The effect of the nanoparticles on the polymers
nano-fillers for multifunctional polymer nanocompos- attached to their surface, and on the bulk polymer
ites (Guo et al. 2008a, 2009a, b; Zhang et al. 2009), itself, has not been systematically investigated. Such
and lithium ion batteries (Wu et al. 2006, 2008). In effects will play an important role in controlling the
addition, the small band gap (*2.1 eV), high-corro- quality and properties of the fabricated polymer
sion resistance, and low cost make iron oxide nanocomposites. In this article, a facile, economic,
nanoparticles suitable for serving as photoelectrodes and solvent-free method utilizing a mild annealing
in solar energy conversions (Ohmori et al. 2000). process was developed to fabricate crystalline iron
Various synthetic methods have been developed oxide nanoparticles. Sodium chloride, which is envi-
for the preparation of hematite nanoparticles includ- ronmentally benign and easy to remove, was used as
ing chemical precipitation (Glisenti 1998), hydro- the template/spacer to prevent the nanoparticles from
thermal reaction (Chen et al. 2002; Dhage et al. 2002; agglomerating. The biodegradable and water soluble
Li et al. 1998, 2007; Wang et al. 2004), sol–gel material, polyvinyl alcohol (PVA), was chosen as the
method (Dong and Zhu 2002; Huo et al. 2005), binding polymer. The ratio of the iron precursor to
solvothermal method (Zheng et al. 2007), ball milling sodium chloride and the presence of PVA in the
process (Bercoff et al. 2007; Wang and Jiang 2007), purifying solution were investigated for particle
and microemulsion technique (Nassar and Husein growth. The effect of the nanoparticles on the phys-
2006; Santra et al. 2001). Iron oxides have strong icochemical properties of PVA was also explored, as
magnetic properties and aggregate/agglomerate spon- the PVA was chemically bound to or physically
taneously, resulting in nanoparticles with a wide size entangled with the nanoparticle surface. The nano-
distribution. In order to prevent nanoparticle agglom- composites consisting of the PVA matrix, reinforced
eration, various surfactants, polymers, or coupling with the fabricated Fe2O3 nanoparticles, were fabri-
agents had been used. Unfortunately, organic sol- cated by the drop casting method. The effects of
vents, bases, and most of the surfactants are harmful nanoparticles on the PVA, both on the nanoparticle
to the environment and expensive for the above surface and on the bulk polymer, were investigated by
reported methods, especially, the wet chemical fab- various analytical methods and discussed.
rications. In addition, most of the traditional methods
are conducted on a small batch scale, and are not
suitable for large-scale fabrication. Thermal decom- Experiment
position of iron nitrate has been investigated by
thermogravimetric analysis (TGA), and a-Fe2O3 was Materials
reported (Wieczorek-Ciurowa and Kozak 1999).
However, there are no reports in the literature Iron nitrite (Fe(NO3)39H2O, Alfa Aesar Company,
regarding the nanostructural formation of a-Fe2O3 404.15 g/mol) and sodium chloride (NaCl, Alfa
from decomposition, especially in the presence of an Aesar Company) are used as the iron precursor and
inert salt. The development of a facile and economic the solid stabilizer to disperse the iron nitrate and
method for large-scale fabrication of hematite nano- prevent the products from agglomerating.
particles still remains a challenge. Polyvinyl alcohol (PVA, MW = 88000–96800,
One of the greatest challenges in fabricating high- degree of polymerization = 2000–2200) was used
quality polymer nanocomposites is the lack of the as the polymer nanocomposite matrix and the stabi-
synergy between the polymer matrices and the lizer to prevent nanoparticle agglomeration. All the
nanofillers. In principle, the polymers could serve chemicals were used as-received without any further
as a template for nanoparticle fabrication. The treatment.
nanoparticles can introduce unique physicochemical
properties into the polymer matrices, such as mag- Nanoparticle fabrication
netism, non-linear optical properties, and improved
mechanical properties. However, the nanoparticle Iron nitrate and sodium chloride were dissolved in de-
materials could also have a deleterious effect on the ionized water with a molar ratio of 1:50 and 1:20,
polymer matrix. respectively. The homogeneous solid solution was

123
J Nanopart Res (2010) 12:2415–2426 2417

obtained by drying at 70 °C under magnetic stirring. sample position on the instrument center. XRD scans
The NPs were manufactured by annealing the solid were recorded from 7 to 77° for 2h with a 0.050°
salt at 700 °C in a tube furnace for about 30 min and step-width and a 60 s counting time for each step.
cooled naturally to room temperature. The product The XRD data were analyzed using the DIFFRAC-
was washed thoroughly with de-ionized water to Plus EVA program (Bruker AXS, Karlsruhe, Ger-
remove sodium chloride and dried completely at many), and the patterns were identified using the
60 °C. In order to prevent particle agglomeration and ICDD PDFMaint computer reference database.
modify the surface chemistry of the NPs, 1 wt% PVA Fourier transform infrared spectroscopy (FT-IR, a
aqueous solution was used to treat the NPs immedi- Bruker Inc. Tensor 27 FT-IR spectrometer with
ately after annealing. The particles were then washed hyperion 1000 ATR microscopy accessory) was used
thoroughly with de-ionized water to remove NaCl. to characterize the surface chemistry of the nanopar-
ticles, as well as to characterize their interaction with
Nanocomposite fabrication the polymer matrix.
Scanning electron microscopy (SEM, JEOL field
The as-produced or functionalized Fe2O3 NPs with emission scanning electron microscope, JSM-6700F)
specific amount were dispersed in 5 wt% PVA was used to investigate the fabricated nanoparticles
aqueous solution (20 mL) for polymer nanocompos- and its dispersion in the polymer matrix. A thin gold
ite fabrication. The aqueous iron oxide nanoparticle layer was deposited to improve the electrical con-
PVA solution was cast in a petri dish and dried ductivity for better imaging.
completely at 60 °C. It was found that the PVA- Magnetic properties of the nanoparticles at room
treated iron oxide nanoparticles have a better disper- temperature were investigated in a 9 T physical
sion quality than those washed with only de-ionized properties measurement system (PPMS) by Quantum
water. Thus, only PVA-treated iron oxide nanoparti- Design.
cles with an iron nitrate to sodium chloride ratio of The differential scanning calorimetry (DSC) tem-
1:20 were used for PVA–Fe2O3 nanocomposite perature and heat flow values were calibrated with an
fabrication and property characterization. indium standard. The heating and cooling rate was
10 °C/min, and the experimental was performed in a
Characterization continuous flow of nitrogen gas with a flow rate of
20 cm3/min (ccpm). The temperature was in the
The thermal decomposition of iron nitrate was range of 25–250 °C.
determined by TGA from 25 to 1000 °C with an air
flow rate of 50 cm3/min (ccpm) and a heating rate of
20 °C/min. Results and discussion
The phase structure of the produced iron oxide
nanoparticles was investigated by powder X-ray Nanoparticle fabrication and property
diffraction. The powder X-ray diffraction analysis characterization
of the samples was carried out with a Bruker AXS D8
Discover diffractometer with General Area Detector Figure 1a and b shows the thermal decomposition
Diffraction System (GADDS) operating with a Cu Ka study of iron nitrate under TGA with the temperature
radiation source filtered with a graphite monochro- changing from 20 to 800 °C, and isothermal decom-
mator (k = 1.5406 Å). The detector used was a HI- position at 700 °C, respectively. The TGA was
STAR two-dimensional multi-wire area detector. The carried out in air with a flow rate of 40 cm3/min.
samples were loaded onto double sided scotch tape, Iron nitrate was observed to decompose in both cases.
placed on a glass slide, and mounted on a quarter- A total of 60.2% was observed to lose sharply due to
circle Eulerian cradle (Huber) on a XYZ stage. The the decomposition of the iron nitrate and additional
X-ray beam was generated at 40 kV and 40 mA 8.0% loss at higher temperatures was due to the loss
power, and was collimated to about 800 lm spot size of the physicochemical adsorption of water. The total
on the sample. The incident x angle was 5°. A laser/ weight residue in both cases is about 19.4 and 19.7%,
video system was used to ensure the alignment of the respectively. This is almost equivalent to the

123
2418 J Nanopart Res (2010) 12:2415–2426

Fig. 1 a Thermal
decomposition study of iron
nitrate and b isothermal
thermal decomposition of
iron nitrate at 700 °C under
air condition

theoretical calculation (19.80%) of iron oxide (a- peak width varies, all the samples have common peak
Fe2O3) from iron nitrate. The residue after each TGA at 24.2°, 33.4°, 35.7°, 41.1°, 49.8°, 54.4°, 62.9°,
run was observed to be stable at 400 °C during the 64.5°, and 72.5°. The d-spacing is calculated using
temperature sweep. The thermal decomposition of Bragg’s law (Rudel and Zite-Ferenczy 1971):
iron nitrate in the isothermal (700 °C) reaction was
nk
completed within 5 min. The iron oxide nanoparticle d¼ ð1Þ
2 sin h
fabrications were carried out at 700 °C with a
reaction time of 1 h. Sodium chloride was chosen where n is chosen as 1 and k is 1.5406 Å for the
as a spacer to disperse iron nitrate and also as a wavelength of Cu Ka radiation.
stabilizer to prevent agglomeration of the formed The calculated d-spacings are 3.67, 2.68, 2.51,
nanoparticles. 2.19, 1.83, 1.68, 1.48, 1.44, and 1.30 Å, respectively.
Figure 2 shows the X-ray diffraction (XRD) These d-spacings are in agreement with those of the
patterns of iron oxide nanoparticles fabricated with XRD PDF card #33-0664 (standard XRD pattern of
different ratios of iron nitrate to sodium chloride (20 a-Fe2O3). This indicates that the formed dark reddish
and 50) and treated with de-ionized water or 1 wt% brown iron oxide nanoparticles are a-Fe2O3. The
PVA aqueous solution, respectively. Although, the corresponding crystal planes are (0, 1, 2), (1, 0, 4),

123
J Nanopart Res (2010) 12:2415–2426 2419

Fig. 2 XRD patterns of


iron oxide nanoparticles
formed (a) with a ratio of
iron nitrate to sodium
chloride (1:20) and washed
with de-ionized water; (b)
with a ratio of iron nitrate to
sodium chloride (1:20) and
treated with PVA aqueous
solution; (c) with a ratio of
iron nitrate to sodium
chloride (1:50) and washed
with de-ionized water; and
(d) with a ratio of iron
nitrate to sodium chloride
(1:50) and washed with
PVA aqueous solution

(1, 1, 0), (1, 1, 3), (0, 2, 4), (1, 1, 6), (2, 1, 4), (3, 0, 0),
and (1, 0, 10), respectively.
The average particle size (L) was estimated from
the Debye–Scherrer equation (Eq. 2; Talapin et al.
2001a):
Kk
L¼ ð2Þ
bð2hÞ cos h
where b(2h) is the full width at half-maximum
(FWHM), K is a constant taken as the normal value
of 0.9, k is the wavelength of X-ray wavelength (for
copper, k = 1.5406 Å), and h is the Bragg angle. The
(1, 1, 0) crystal plane at 2h = 35.7° was used to
estimate the particle size. The calculated values are
about 19, 15, 14 and 13 nm, respectively. The
particle sizes are observed to decrease with an
Fig. 3 FT-IR spectra of iron oxide nanoparticles formed (a)
increase in the ratio of sodium chloride to iron with a ratio of iron nitrate to sodium chloride (1:20) and
nitrite. The solution containing PVA for salt removal washed with de-ionized water; (b) with a ratio of iron nitrate to
was also observed to inhibit the particle growth by sodium chloride (1:20) and treated with PVA aqueous solution;
serving as a stabilizer. The presence of PVA has (c) with a ratio of iron nitrate to sodium chloride (1:50) and
washed with de-ionized water; and (d) with a ratio of iron
effectively prevented the Ostwald ripening process nitrate to sodium chloride (1:50) and washed with PVA
(Guo et al. 2006a; Talapin et al. 2001b) and favors aqueous solution
smaller nanoparticle fabrication.
Figure 3 shows the FT-IR spectra of the as-
produced a-Fe2O3 nanoparticles and those function- in the nanoparticles washed with PVA aqueous
alized with PVA. The characteristic hydrogen-bonded solution. The absorption band at 1,142 cm-1 has
stretch bands (3,259 and 1,645 cm-1; Matveev et al. been used as an assessment tool of the PVA structure
2005; Okuno et al. 2003; Ping et al. 2001), C–H due to the semi-crystalline nature of the synthetic
stretch bands (1,373 and 1,319 cm-1; Carrillo et al. polymer (Mansur et al. 2004). Thus, the 1,142 cm-1
2004), and C–O bands (1,142 and 1,088 cm-1; vibration band observed in the spectra of the nano-
Carrillo et al. 2004; Okuno et al. 2003) are observed particles treated with PVA aqueous solution, clearly

123
2420 J Nanopart Res (2010) 12:2415–2426

indicates the existence of PVA. The absorption bands characterized with a differential scanning calorimeter
at 2,949 and 2,904 cm-1 represent the sp3 bonding in (DSC), which is sensitive to the thermal history of the
–CH3 group.(Deschenaux et al. 1999) The peaks samples. The first heating and cooling run of the DSC
around 831 cm-1 are attributed to C–H bending at a rate of 10 K/min and a nitrogen flow rate of
(Mansur et al. 2004). No C–H group peak was 10 ccpm was used to characterize the thermal prop-
observed in the spectra of the nanoparticles washed erties of all samples. The heating run was used to
with de-ionized water, and the obvious hydroxyl obtain the melting temperature and melting enthalpy,
groups indicate the presence of physically absorbed while the cooling run was used to characterize the
moisture or chemically hydrolyzed water, which crystallization and fusion energy. The results are
make the nanoparticles compatible with PVA, and presented in Fig. 5 and summarized in Table 1. The
are responsible for the subsequent uniform particle thermograms show two endothermic peaks. The first
dispersion in the PVA nanocomposite fabrication. one at lower temperature is due to the melting of the
Figure 4 shows the TGA curves of the nanopar- crystallites of the cross-linked network, and the
ticles treated with de-ionized water and with the PVA second peak in the range of 200–240 °C is attributed
aqueous solution, respectively. The small weight to the melting of the PVA upon recrystallization
percentage loss in the iron oxide nanoparticles is due (Agrawal and Awadhia 2004). The melting temper-
to the evaporation of the moisture or the condensation ature arising from cross-linking decreased from
of the hydroxyl groups. The initial decomposition 124 °C (pure PVA) to 88 °C (PVA attached to
temperature (sharp weight loss) was observed to nanoparticles fabricated with a ratio of 1:20) and
increase from 233 to 302 °C, when the ratio of iron further down to 80 °C (PVA attached to nanoparticles
nitrite to sodium chloride increased from 20 to 50, fabricated with a ratio of 1:50). This observation
which is due to the smaller particle size in the latter indicates that the presence of the iron oxide nano-
case. The stronger interaction between the iron oxide particles prohibits the extent of the cross-linking
nanoparticles and the PVA matrix, arising from the during the formation of solid PVA from the aqueous
higher specific surface area inherent with the smaller polymer solution. The melting temperature upon
nanoparticles improves the thermal stability of the recrystallization decreased from 227.5 to 224.2 °C
entangled polymer chains. after the PVA powders were processed into a thin
The interaction between the nanoparticles and the film structure by the aqueous drop casting. The
attached polymer, i.e., the effect of the nanoparticles melting temperature of the PVA increased from 224.2
on the thermal properties of the PVA, was further to 229 °C and 231 °C after it was attached to the iron

Fig. 4 TGA curves of (a)


pure PVA, and iron oxide
nanoparticles washed with
(a) water, and (c) 1 wt%
PVA, respectively (ratio of
iron nitrate to sodium
chloride 1:20)

123
J Nanopart Res (2010) 12:2415–2426 2421

Fig. 5 DSC thermograms


of (a) pure PVA thin film,
and PVA attached to iron
oxide nanoparticles
fabricated with a ratio of
iron nitrite to sodium
chloride of (b) 1:20 and
(c)1:50, respectively

Table 1 Thermal properties and calculated crystallinity of pure PVA and bounded PVA
Tm (°C) Tc (°C) DHm (J/g) DHf (J/g) Crystallinity (%)

PVA powder 227.5 198.6 52.43 43.08 37.8


PVA film 224.2 201.7 38.93 41.11 28.1
a
PVA-NPs 229.0 204.6 85.75 49.69 35.8
PVA-NPsb 231.3 191.4 113.3 53.44 38.5
a
PVA on the surface of the nanoparticle surface, fabricated with a ratio of 1:20 (Na:Fe)
b
PVA on the surface of the nanoparticle surface, fabricated with a ratio of 1:50 (Na:Fe)

oxide nanoparticle surface fabricated with a ratio of oxide nanoparticles. The nanoparticle size was also
iron nitrate to sodium chloride of 1:20 and 1:50, observed to have a significant effect on the melting
respectively. This observation is consistent with the temperature and the melting enthalpy.
enhanced thermal stability with a higher initial The degree of the relative crystallinity (Xc; Kim
decomposition temperature (Fig. 5), and it is different et al. 2008; Peppas and Merrill 1976) of the PVA
from the PVA in the hydroxyapatite (Hap)–PVA attached on the nanoparticle surface was estimated
composite system (Kim et al. 2008). from the endothermic area (from cooling stage) using
The melting enthalpy (DHm, based on the weight Eq. 3:
of pure polymer rather than the total weight of DHf
polymer nanocomposite) was observed to increase Xc ¼ ð3Þ
DHf0
significantly from 38.93 to 85.75 J/g and 113.3 J/g,
which is consistent with the nanocomposites com- where DHf is the measured enthalpy of fusion from
prised of silver nanoparticles reinforced with PVA the DSC thermogram (based on the weight of the
(Mbhele et al. 2003). This is due to the reduced pure polymer rather than the total weight of the
mobility of the polymer chains after being attached to polymer nanocomposites), and DHf0 is the enthalpy of
the nanoparticle surface. The dramatic difference in fusion of 100% crystalline PVA (DHm is 138.6 J/g;
melting temperature between the pure PVA and the Peppas and Merrill 1976). The crystallinity of PVA
PVA attached to the nanoparticle surface indicates a decreased after the PVA powder was drop-casted into
strong interaction between the PVA and the iron the thin film structure. However, the crystallinity

123
2422 J Nanopart Res (2010) 12:2415–2426

Fig. 6 Magnetic hysteresis


loop (M–H) of iron oxide
nanoparticles fabricated
with ratios of iron nitrate to
sodium chloride (1:50 and
1:20); left inset shows the
enlarged hysteresis loop at
low field, the right inset
shows the magnetization
over H-1 for Ms calculation

increased after PVA physicochemically bonded to the Polymer nanocomposites


nanoparticle surface.
Figure 6 shows the magnetic properties of the Figure 7 shows the SEM microstructure of the PVA
nanoparticles fabricated with different ratios of iron nanocomposites reinforced with a particle loading of
nitrate to sodium chloride. The saturation magneti- 10 wt%. The nanoparticles were observed to be
zation (Ms) was not reached even at a high field of 9 uniformly dispersed in the polymer matrix without
Tesla. Ms was determined by the extrapolated satu- obvious agglomeration. No cracks or voids were
ration magnetization obtained from the intercept of evident in the polymer nanocomposites. The high-
magnetization versus H-1 at high field (Chen et al. resolution SEM image (inset of Fig. 7) shows an
1994; Guo et al. 2006a). The calculated saturation interface between a nanoparticle and the polymer
magnetization of the nanoparticles fabricated with the matrix, which is similar to that observed in nano-
ratio of iron nitrate to sodium chloride (1:50 and composites of alumina nanoparticles reinforced vinyl
1:20) was 3.3 and 3.0 emu/g, respectively. The ester resin (Guo et al. 2006b). The fabricated
coercivity (coercive force, Hc) is 350 and 300 Oe nanoparticles are surrounded by hydroxyl groups,
for the nanoparticles fabricated with the ratio of iron which form hydrogen bonding with the PVA, and
nitrate to sodium chloride 1:20 and 1:50, respec- thus, lead to void- and crack-free structural polymer
tively. The difference is due to the particle size effect. nanocomposites. These strong hydrogen bonds are
The coercivity is smaller than that (784.5 Oe) of the responsible for the strong interaction between the
a-Fe2O3 nanoparticles dispersed in a polystyrene nanoparticles and the polymer matrix.
matrix (Zhang et al. 1989) and that (2,279.0 Oe) of Figure 8 shows the TGA curves of the pure PVA
the cantaloupe-like Fe2O3 (Zhu et al. 2007). This and its nanocomposites with a nanoparticle loading of
could be due to the morphological (shape and size) 1, 6, and 10 wt%, respectively. The weight loss at
difference (Zheng et al. 2007). The saturation mag- lower temperature (lower than 100 °C for pure PVA
netization (Ms) is much lower than that of c-Fe2O3 and 116 °C for PVA reinforced with iron oxide
(74 emu/g; Guo et al. 2008a), and that of metallic nanoparticles) is due to the evaporation of adsorbed
iron (218 emu/g; Guo et al. 2008b). However, it is moisture. The loss at higher temperature is due to the
similar to the Ms observed in other nanoparticles with dehydration of the hydrogen bonding between the
different morphologies (Zheng et al. 2007; Zhu et al. PVA and the hydroxy groups in the nanoparticles.
2007). The weight loss in the medium temperature range

123
J Nanopart Res (2010) 12:2415–2426 2423

Fig. 9 XRD patterns of pure (a) PVA and PVA nanocompos-


ites with an iron oxide nanoparticle loading of (b) 1 wt%, (c) 6
wt%, and (d) 10 wt%

C and D was observed in the polymer nanocompos-


ites, which is due to the degradation and carboniza-
tion of the PVA. The formed carbon further reacted
with iron oxide. In other words, the addition of the
iron oxide nanoparticles favors the thermal decom-
Fig. 7 SEM microstructures of the polymer nanocomposites position of the PVA. A similar results is observed
with a particle loading of 10 wt% (the inset shows the high
resolution of SEM)
with CuO nanoparticles (Guo et al. 2007) and Fe2O3
nanoparticles (Guo et al. 2008a), but different from
Al2O3 nanoparticles (Guo et al. 2006b) reinforced
vinyl ester resin nanocomposites.
Figure 9 shows the XRD patterns of the pure PVA
and its nanocomposites reinforced with different iron
oxide nanoparticle loadings. The peaks at 19.5° and
40.5° were observed in the pure PVA. The first peak
has a d-spacing of 0.454 nm, corresponding to the
typical doublet reflection of the (1, 0, 1) and (1, 0,
-1) planes of the semicrystalline atactic PVA. The
second peak is assigned to the (2, 2, 0) plane of the
PVA (Kim et al. 2008; Ricciardi et al. 2004). The
peak corresponding to the (1, 0, 1) plane of the PVA
becomes narrower and narrower as a function of the
particle loading. This indicates that the crystallite
sizes along the (1, 0, 1) lattice direction, as dictated
by Eq. 2, became larger with the addition of more
Fig. 8 TGA curves of pure PVA and PVA nanocomposites
with nanoparticle loading of 1, 6, and 10 wt%, respectively iron oxide nanoparticles. This suggests a higher
crystallinity of the PVA, which is in contrast to
hydroxyapatite (Hap) nanoparticles reinforced with
(120–247 °C) is due to the condensation of the PVA (Kim et al. 2008). The presence of iron oxide
hydroxyl groups. The sharp weight loss at tempera- nanoparticles favors the recrystallization of PVA
tures higher than 247 °C is due to the thermal during the polymer solution solidification process.
decomposition of the PVA. A plateau in the range of Concurrently, the standard Fe2O3 reflection peaks are

123
2424 J Nanopart Res (2010) 12:2415–2426

observed, similar to Fig. 2. The difference in XRD loading, which is in stark contrast to the results when
spectra between the iron oxide nanoparticles and the the PVA is attached to the nanoparticle surface
polymer nanocomposites is evident in Fig. 9. The (Fig. 5, Table 1). This behavior is similar to what
iron oxide peak became broader, indicating a uniform was observed in nanocomposites comprised of silver
dispersion of nanoparticles, consistent with the SEM nanoparticles embedded in PVA (Mbhele et al. 2003).
observation (Fig. 8).
In order to investigate the effect of nanoparticle
loading on the thermal properties of the bulk polymer Conclusions
matrices, such as melting and crystallization (forma-
tion of the crystalline structure), DSC measurements An environmentally benign method was developed to
were run, and the results are presented in Fig. 10 and fabricate iron oxide (a-phase) nanoparticles. Inexpen-
summarized in Table 2. The lower melting temper- sive sodium chloride effectively serves as a solid
ature (123 °C) of the PVA nanocomposites with spacer to disperse the salt and prevent the iron oxide
1 wt% nanoparticle loading, as compared to the pure nanoparticles from agglomerating. The nanoparticle
PVA thin film, decreased slightly and then increased size increased with the increase of the ratio of iron
to 161.5 °C for a particle loading of 6 wt% and nitrate to sodium chloride. The presence of PVA in the
151.9 °C for a particle loading of 10 wt%. The higher solution limits the particle size and favors a more
melting temperature decreased with the increase in uniform particle-dispersed polymer nanocomposite
nanoparticle loading. Similarly, the melting temper- fabrication. Obvious changes of structural phase and
ature and the melting enthalpy of the PVA matrices thermal properties, such as low-temperature cross-
decreased with the increase of the nanoparticle linking, high-temperature recrystallization, melting
enthalpy, fusion enthalpy, and crystallinity have been
observed once the PVA was attached to the nanopar-
ticle surface. The presence of the nanoparticles
increased the cross-linking with an increase in the
lower melting temperature. The thermal stability was
enhanced when PVA was attached to the nanoparticle
surface. The nanoparticles fabricated showed that a
coercivity of 350 Oe and a lower saturation magneti-
zation. The nanoparticles were uniformly dispersed in
the polymer matrix as shown in the SEM microstruc-
ture analysis and the XRD pattern variation. SEM
images also indicated an intermediate phase, suggest-
ing a strong interaction between the nanoparticles and
the polymer matrix. In summary, the addition of
nanoparticles favored the cross-linkage of the PVA,
resulting in an increase in the lower temperature
Fig. 10 DSC thermograms of (a) pure PVA thin film, PVA
nanocomposites with a particle loading of (b) 1 wt%, (c) 6 melting point, and thus a subsequent improvement in
wt%, and (d) 10 wt% (Exo, up) the thermal stability of the nanocomposite.

Table 2 Thermal properties


Tm (°C) Tc (°C) DHm (J/g) DHf (J/g) Crystallinity
and calculated crystallinity of
(%)
pure PVA and polymer
nanocomposites PVA powder 227.5 198.6 52.43 43.08 37.8
Pure PVA film 224.2 201.7 38.93 41.11 28.1
1 wt% 227.3 203.0 43.16 44.72 32.3
6 wt% 223.6 201.1 31.91 44.55 32.1
10 wt% 221.8 193.8 23.53 39.27 28.3

123
J Nanopart Res (2010) 12:2415–2426 2425

Acknowledgments The project is partially supported by the Guo Z, Lei K, Li Y, Ng HW, Hahn HT (2008a) Fabrication and
start-up grant and research enhancement grant from Lamar characterization of iron oxide nanoparticle reinforced
University. The authors kindly acknowledge the support from vinyl-ester resin nanocomposites. Compos Sci Technol
Northrop–Grumman Corporation. DPY acknowledges support 68:1513–1520
from the NSF under Grant No. DMR 04-49022. FT-IR analysis Guo Z, Lin H, Karki AB, Young DP, Hahn HT (2008b) Facile
done by Dr. Y. Mou from Department of Chemistry and monomer stabilization approach to fabricate iron/vinyl
Physics at LU and financial support from the Welch ester resin nanocomposites. Compos Sci Technol
Foundation (V-1103) are kindly acknowledged. 68:2551–2556
Guo Z, Shin K, Karki AB, Young DP, Hahn HT (2009a)
Fabrication and characterization of iron oxide nanoparti-
cles reinforced polypyrrole nanocomposites. J Nanopart
References Res 11:1441–1453
Guo Z, Lee SE, Kim H, Park S, Hahn HT, Karki AB, Young
Agrawal SL, Awadhia A (2004) DSC and conductivity studies DP (2009b) Fabrication, characterization and microwave
on PVA based proton conducting gel electrolytes. Bull properties of polyurethane nanocomposites reinforced
Mater Sci 27:523–527 with iron oxide and barium titanate nanoparticles. Acta
Bercoff PG, Bertorello HR, Oliva MI (2007) Memory effect of Mater 57:267–277
ball-milled and annealed nanosized hematite. Physica B Huo L, Li Q, Zhao H, Yu L, Gao S, Zhao J (2005) Sol–gel
398:204–207 route to pseudocubic shaped a-Fe2O3 alcohol sensor:
Carrillo F, Colom X, Sunol JJ, Saurina J (2004) Structural preparation and characterization. Sens Actuat B 107:915–
FTIR analysis and thermal characterisation of lyocell and 920
viscose-type fibers. Eur Polym J 40:2229–2234 Kay A, Cesar I, Gratzel M (2006) New benchmark for water
Chen JP, Sorensen CM, Klabunde KJ, Hadjipanayis GC (1994) photooxidation by nanostructured a-Fe2O3 films. J Am
Magnetic properties of nanophase cobalt particles syn- Chem Soc 128:15714–15721
thesized in inversed micelles. J Appl Phys 76:6316–6318 Kim G-M, Asran AS, Michler GH, Simon P, Kim J-S (2008)
Chen LX, Liu T, Thurnauer MC, Csencsits R, Rajh T (2002) Electrospun PVA/HAp nanocomposite nanofibers: bi-
Fe2O3 nanoparticle structures investigated by X-ray omimetics of mineralized hard tissues at a lower level of
absorption near-edge structure, surface modifications, and complexity. Bioinspir Biomim 3:046003
model calculations. J Phys Chem B 106:8539–8546 Li Y, Liao H, Qian Y (1998) Hydrothermal synthesis of
Deschenaux C, Affolter A, Magni D, Hollenstein C, Fayet P ultrafine a-Fe2O3 and Fe3O4 powders. Mater Res Bull
(1999) Investigations of CH4, C2H2 and C2H4 dusty RF 33:841–844
plasmas by means of FTIR absorption spectroscopy and Li L, Chu Y, Liu Y, Dong L (2007) Template-free synthesis
mass spectrometry. J Phys D 32:1876 and photocatalytic properties of novel Fe2O3 hollow
Dhage SR, Khollam YB, Potdar HS, Deshpande SB, Bakare spheres. J Phys Chem C 111:2123–2127
PP, Sainkar SR, Date SK (2002) Effect of variation of Mansur HS, Orefice RL, Mansur AAP (2004) Characterization
molar ratio (pH) on the crystallization of iron oxide of poly(vinyl alcohol)/poly(ethylene glycol) hydrogels
phases in microwave hydrothermal synthesis. Mater Lett and PVA-derived hybrids by small-angle X-ray scattering
57:457–462 and FTIR spectroscopy. Polymer 45:7193–7202
Dong W, Zhu C (2002) Use of ethylene oxide in the sol-gel Matveev S, Portnyagin M, Ballhaus C, Brooker R, Geiger CA
synthesis of a-Fe2O3 nanoparticles from Fe(III) salts. (2005) FTIR spectrum of phenocryst olivine as an indi-
J Mater Chem 12:1676–1683 cator of silica saturation magmas. J Petrol 46:603–614
Feldmann C (2001) Preparation of nanoscale pigment particles. Mbhele ZH, Salemane MG, van Sittert CGCE, Nedeljkovic
Adv Mater 13:1301–1303 JM, Djokovic V, Luyt AS (2003) Fabrication and char-
Fukazawa M, Matuzaki H, Hara K (1993) Humidity- and gas- acterization of silver–polyvinyl alcohol nanocomposites.
sensing properties with an Fe2O3 film sputtered on a Chem Mater 15:5019–5024
porous Al2O3 film. Sens Actuat B14:521–522 Nassar N, Husein M (2006) Preparation of iron oxide nano-
Glisenti A (1998) Interaction of formic acid with Fe2O3 pow- particles from FeCl3 solid powder using microemulsions.
ders under different atmospheres: an XPS and FTIR study. Phys Status Solidi A 203:1324–1328
J Chem Soc Faraday Trans 94:3671–3676 Neri G, Bonavita A, Galvagno S, Siciliano P, Capone S (2002)
Guo Z, Henry LL, Palshin V, Podlaha EJ (2006a) Synthesis of CO and NO2 sensing properties of doped-Fe2O3 thin films
poly(methyl methacrylate) stabilized colloidal zero- prepared by LPD. Sens Actuat B 82:40–47
valence metallic nanoparticles. J Mater Chem 16:1772– Peppas NA, Merrill EW (1976) Differential scanning calo-
1777 rimetry of crystallized PVA hydrogels. J Appl Polym Sci
Guo Z, Tony P, Oyoung C, Wang Y, Hahn HT (2006b) Surface 20:1457–1465
functionalized alumina nanoparticle filled polymer nano- Ohmori T, Takahashi H, Mametsuka H, Suzuki E (2000)
composites with enhanced mechanical properties. J Mater Photocatalytic oxygen evolution on a-Fe2O3 films using
Chem 16:2800–2808 Fe3? ion as a sacrificial oxidizing agent. Phys Chem
Guo Z, Liang X, Pereira T, Scaffaro R, Hahn HT (2007) CuO Chem Phys 2:3519–3522
nanoparticle reinforced vinyl-ester resin nanocomposites: Okuno D, Lwase T, Shinzawa-Itoh K, Yoshikawa S, Kitagawa
fabrication, characterization and property analysis. Com- T (2003) FTIR detection of protonation/deprotonation of
pos Sci Technol 67:2036–2044 key carboxyl side chains caused by redox change of the

123
2426 J Nanopart Res (2010) 12:2415–2426

CuA-Heme a moiety and ligand dissociation from the Wang Y, Cao J, Wang S, Guo X, Zhang J, Xia H, Zhang S, Wu
Heme a3-CuB center of bovine heart cytochrome c oxi- S (2008) Facile synthesis of porous a-Fe2O3 nanorods and
dase. J Am Chem Soc 125:7209–7218 their application in ethanol sensors. J Phys Chem C
Ping ZH, Nguyen QT, Chen SM, Zhou JQ, Ding YD (2001) 112:17804–17808
States of water in different hydrophilic polymers: DSC Wieczorek-Ciurowa K, Kozak AJ (1999) The thermal
and FTIR studies. Polymer 42:8461–8467 decomposition of Fe(NO3)39H2O. J Therm Anal Calorim
Ren Y, Ma Z, Qian L, Dai S, He H, Bruce PG (2009) Ordered 58:647–651
crystalline mesoporous oxides as catalysts for CO oxida- Wu C, Yin P, Zhu X, OuYang C, Xie Y (2006) Synthesis of
tion. Catal Lett 131:146–154 hematite (a-Fe2O3) nanorods: diameter-size and shape
Ricciardi R, Auriemma F, De Rosa C, Laupretre F (2004) effects on their applications in magnetism, lithium ion
X-ray diffraction analysis of Poly(vinyl alcohol) hydro- battery, and gas sensors. J Phys Chem B 110:17806–17812
gels, obtained by freezing and thawing techniques. Mac- Wu X-L, Guo Y-G, Wan L-J, Hu C-W (2008) A-Fe2O3
romolecules 37:1921–1927 nanostructures: inorganic salt-controlled synthesis and
Rofer-Depoorter CK (1981) A comprehensive mechanism for their electrochemical performance toward lithium storage.
the Fischer-Tropsch synthesis. Chem Rev 81:447–474 J Phys Chem C 112:16824–26829
Rudel R, Zite-Ferenczy F (1971) Interpretation of Light Dif- Xie X, Li Y, Liu Z-Q, Haruta M, Shen W (2009) Low-tem-
fraction By cross-striated muscle as Bragg reflexion of perature oxidation of CO catalysed by Co3O4 nanorods.
light by the lattice of contractile proteins. J Physiol Nature 458:746–749
290:317–330 Yu B, Zhu C, Gan F (2000) Large nonlinear optical properties
Santra S, Tapec R, Theodoropoulou N, Dobson J, Hebard A, of Fe2O3 nanoparticles. Physica E 8:360–364
Tan W (2001) Synthesis and characterization of silica- Zhang D, Karki AB, Rutman D, Young DP, Wang A, Cocke D,
coated iron oxide nanoparticles in microemulsion: the Ho TH, Guo Z (2009) Electrospun polyacrylonitrile
effect of nonionic surfactants. Langmuir 17:2900–2906 nanocomposite fibers reinforced with Fe3O4 nanoparti-
Talapin DV, Haubold S, Rogach AL, Kornowski A, Haase M, cles: fabrication and property analysis. Polymer 50:4189–
Weller H (2001a) A novel organometallic synthesis of 4198
highly luminescent CdTe nanocrystals. J Phys Chem B Zhang Y, Chu Y, Dong L (1989) One-step synthesis and
105:2260–2263 properties of urchin-like PS/a-Fe2O3 composite hollow
Talapin DV, Rogach AL, Haase M, Weller H (2001b) Evolu- microspheres. Nanotechnology 18:435608
tion of an ensemble of nanoparticles in a colloidal solu- Zheng Y, Cheng Y, Wang Y, Bao F, Zhou L, Wei X, Zhang Y,
tion: theoretical Study. J Phys Chem B 105:12278–12285 Zheng Q (2007) Quasicubic a-Fe2O3 nanoparticles with
Wang L-L, Jiang J-S (2007) Preparation of a-Fe2O3 nanopar- excellent catalytic performance. J Phys Chem B 110:
ticles by high-energy ball milling. Physica B 390:23–27 3093–3097
Wang X, Chen X, Ma X, Zheng H, Ji M, Zhang Z (2004) Low- Zhu L-P, Xiao H-M, Fu S-Y (2007) Template-free synthesis of
temperature synthesis of a-Fe2O3 nanoparticles with a monodispersed and single-crystalline cantaloupe-like
closed cage structure. Chem Phys Lett 384:391–393 Fe2O3 superstructures. Cryst Growth Des 7:177–182

123

You might also like