Release of Nanoclay and Surfactant From Polymer Clay Nanocomposites Into A Food Simulant

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

pubs.acs.org/est

Release of Nanoclay and Surfactant from Polymer−Clay


Nanocomposites into a Food Simulant
Yining Xia, Maria Rubino,* and Rafael Auras
School of Packaging, Michigan State University, East Lansing, Michigan 48824-1226, United States
*
S Supporting Information

ABSTRACT: Release assessment of organo-modified mont-


morillonite (O-MMT) nanoclay and the organo-modifiers
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

(surfactants) was performed on two types of polymer−clay


nanocomposites: polypropylene (PP) and polyamide 6 (PA6)
Downloaded via UNIV OF NEW ENGLAND on January 11, 2019 at 07:31:32 (UTC).

with O-MMT. In accordance with ASTM D4754-11, nano-


composite films were exposed to ethanol as a fatty-food
simulant at 70 °C. The release of O-MMT, with Si and Al used
as the nanoclay markers, was evaluated by graphite furnace
atomic absorption spectrometry. The nanoclay particles
released in ethanol were visualized by transmission electron
microscopy (TEM). More nanoclay particles were released
from PP−clay films (0.15 mg L−1) than from PA6−clay films
(0.10 mg L−1), possibly due to the lack of interaction between
the nanoclay and PP as indicated by the structure and morphology in the TEM images. The surfactant release was quantified by a
liquid chromatography tandem mass spectrometry (LC-MS/MS) method. A substantial amount of surfactant was released into
ethanol (3.5 mg L−1 from PP−clay films and 16.2 mg L−1 from PA6−clay films), indicating changes in the nanoclay structure
within the nanocomposite while it was exposed to ethanol. This research has provided information for the determination of
exposure doses of nanoclay and surfactant in biosystems and the environment, which enabled the risk assessment.

■ INTRODUCTION
The use of nanocomposites consisting of polymers and
cationic surfactants (e.g., alkylammonium cations), to improve
compatibility of the nanoclay with the polymer.10 Nanoclays are
engineered nanoparticles (ENPs) is expanding rapidly, with added to several polymer matrices including polypropylene and
global sales of over US$1.2 billion in 2013 rising to an low density polyethylene to improve their barrier (e.g., to water
estimated US$4.2 billion (U.S.) by 2019.1 The addition of vapor and gases such as oxygen and carbon dioxide) and
ENPs at small loadings significantly improves the performance mechanical properties.6,11 By adding nanoclay, thinner films can
of polymer materials and therefore expands their applications. be produced having similar strength and barrier properties as
For example, the use of nanoscale metals enhances the thicker films without nanoclay, and solid waste can be reduced.
antimicrobial activity and UV resistance of polymers;2,3 the For novel biobased plastics, such as poly(lactic acid) and
incorporation of carbon nanotubes improves thermal, mechan- thermoplastic starch, the incorporation of nanoclay has
ical, and electrical properties of polymers;4,5 and the addition of expanded the range of applications of these materials by
nanoclays increases the barrier properties and heat stability of overcoming their performance limitations (e.g., low barrier to
polymers.6,7 moisture, low heat-deflection temperature).9,12
Nanoclays, such as organo-modified montmorillonite (O- The cytotoxicity of nanoclays has been evaluated in vitro and
MMT), are ENPs increasingly being used in consumer goods in vivo using different cell models such as human epithelial
due to their low cost, commercial availability, high stability, and cells,13 human normal intestinal cells,14 and human hepatic
relatively simple processing. Nanocomposites with O-MMT as cells.15 Studies have shown that nanoclays tend to penetrate
the nanofiller account for over half of total nanocomposite into cells and may affect cell function. Yamashita et al.16
consumption, with the main applications in the automotive demonstrated that nanosilica particles with diameters of less
parts and packaging industries.1,8 MMT is obtained from than 100 nm penetrated and induced structural and functional
naturally occurring layered silicate minerals with a crystal abnormalities in mouse placenta and caused fetal growth
structure consisting of two silica tetrahedral sheets fused to an restriction. Verma et al.13 found that the shape and surface area
edge-shared alumina octahedral sheet.9 The clay layers are
usually parallel stacked to form tactoids with about 1 nm Received: May 29, 2014
interlayer space (or clay gallery) containing exchangeable Revised: September 16, 2014
cations (e.g., Na+ or K+). Organo-modification of MMT is Accepted: November 4, 2014
carried out by replacing the exchangeable cations with organic Published: November 4, 2014

© 2014 American Chemical Society 13617 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624
Environmental Science & Technology Article

of nanoclays impact cell viability; platelet nanoclays were more polymer.30 PA6 resin (Ultramid B40 01) was obtained from
cytotoxic than tubular ones. Also, the potential risks of BASF (Florham Park, NJ, USA).
surfactants used as organo-modifiers of nanoclays have been Two types of nanoclay were used: Nanomer I.44P (herein
investigated, revealing that some surfactants are toxic to referred to as I44P clay) was obtained from Nanocor
ecosystems, animals, and humans.17−19 Furthermore, the (Aberdeen, MS, USA) containing 65 wt % MMT and 35 wt
degradation products of phenyl-containing surfactants may % surfactant (Arquad 2HT-75), and Cloisite 93A (herein
cause endocrine disruption in wildlife or humans.20,21 referred to as Cloisite clay) was obtained from Southern Clay
Nanoparticles including nanoclay may reach biological Products (Gonzales, TX, USA) containing 60 wt % MMT and
systems through different routes (Figure 1). One route of 40% surfactant (Armeen M2HT). The surfactants used in the
nanoclays were also obtained separately (from AkzoNobel, IL,
USA): dimethyl dihydrogenated tallow amine or Arquad 2HT-
75 (used in I44P clay), around 75 wt % purity; and methyl
dihydrogenated tallow amine or Armeen M2HT (used in
Cloisite clay), around 90 wt % purity. Both surfactants consist
of two alkyl chains (hydrogenated tallow) ranging from 12 to
18 carbons with mainly C16 and C18 (>96 wt %) according to
the data sheet from the supplier. Therefore, only three main
components of each surfactant were considered for analysis and
designated as C16C16-, C16C18-, and C18C18-Arquad or Armeen.
Preparation of Polymer−Clay Films. The PP−clay
nanocomposite was prepared by initially mixing the PP and
Figure 1. Routes of potential nanoparticle exposure to the
MAPP resins for 2 min and then melting in a Haake Rheomix
environment and humans.
lab mixer (Thermo Electron Co., Newington, NH, USA) at 180
°C and 40 rpm for 5 min. I44P clay was added and further
mixing was carried out at 180 °C and 150 rpm for 5 min. The
exposure could occur when nanocomposites are used as final composition of the PP−clay nanocomposite was 85 wt %
packaging materials in contact with food;22,23 nanoparticles PP, 12 wt % MAPP and 3 wt % nanoclay. The PA6−clay
may be released from the packaging material into the food. nanocomposite was prepared by predrying the polymer resin in
Other routes of exposure could occur when nanocomposites a vacuum oven at 100 °C for 8 h and then melting in a Haake
are used in manufacturing or buried in landfills; nanoparticles Rheomix lab mixer at 240 °C and 40 rpm for 5 min. Cloisite
may be released into the surrounding environment and reach clay was added and further mixing was carried out at 240 °C
plants, wildlife, or humans.24,25 and 150 rpm for 5 min. The final composition of the PA6−clay
The health and safety risks of nanoparticles on humans and nanocomposite was 95 wt % PA6 and 5 wt % nanoclay. The
the environment are not well understood due to the difficulty in type of clay used with each polymer, as well as the composition
detecting, measuring, and characterizing nanoparticles in of the nanocomposite, was selected to simulate the
different media as well as in evaluating exposure levels to commercialized product.
nanoparticles.26−28 To date, release assessments of nanoclay The polymer−clay nanocomposites prepared by melt mixing
and associated surfactants from nanocomposites are scarce. were ground into small pellets with a Dynisco grinder (Dynisco
Gaining knowledge on the transport of these components from Instruments LLC, Franklin, MA, USA) and further converted
nanocomposites when exposed to different conditions is critical into films. PP−clay films of 22.5 ± 1.1 μm thickness were
to the evaluation of exposure dose and related risk assess- produced by a Killion blown film extruder (Model KL-100,
ment.29 screw size of 1 in., L/D ratio of 30:1; Davis-Standard Corp.,
The aim of this study was to investigate the release of Cedar Grove, NJ, USA) with a temperature profile of 193−204
nanoclay and surfactant from polymer−clay nanocomposite °C (380−400 °F) for the extruder and a screw speed of 14 rpm.
systems into a food simulant by tracking the nanoclay and PA6−clay films of 21.1 ± 1.2 μm thickness were produced by a
surfactant and then correlating the release to interactions Randcastle cast film extruder (Model RCP-0625, screw size of
among the polymer, the nanoclay and the solvent. Such 0.625 in., L/D ratio of 24:1; Extrusion Systems Inc., Cedar
interactions could be intercalation, exfoliation, or aggregation of Grove, NJ, USA) with a temperature profile of 238−246 °C
the nanoclay within the polymer and the effect of the solvent (460−475 °F) for the extruder and a screw speed of 45 rpm.
either on the swelling of the polymer or on the change of clay Control films without nanoclay (22.8 ± 1.8 μm for PP+MAPP,
galleries. Two polymers that are well documented in the and 22.1 ± 1.6 μm for PA6) were also prepared in the same
literature were selected as model systems: polypropylene (PP) manner.
and polyamide 6 (PA6). These polymers differ in polarity and Characterization of Polymer−Clay Films. Thermal
chemical composition and, therefore, represent two different properties of both nanocomposite and control films were
groups of polymers.


characterized by differential scanning calorimetry (DSC) and
dynamic mechanical analysis (DMA). The melting temperature
MATERIALS AND METHODS (Tm) and crystallinity (Xc) were determined with a Q-100 DSC
Materials. PP resin (Profax 6523) was supplied by (TA Instruments, Inc., New Castle, DE, USA) in the first
LyondellBasell Industries (Houston, TX, USA). Maleic heating cycle from 40 to 260 °C at a ramp rate of 10 °C min−1.
anhydride-graf t-polypropylene resin (MAPP or Bondyram The glass transition temperature (Tg) was recorded by a Q-800
1001, 1 wt % bound maleic anhydride) was obtained from DMA (TA Instruments, Inc.) scanning 5 °C min−1 from −50 to
Polyram Co. (Shelby Township, MI, USA) and used to 100 °C for PP-based samples and 0 to 120 °C for PA6-based
improve the compatibility between the nanoclay and the samples. The measurement was carried out in a tension mode
13618 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624
Environmental Science & Technology Article

with a constant strain of 0.1%, a constant frequency of 1 Hz, Before injection into the GFAA spectrometer, the extraction
and a preload of 0.1 N. All samples were tested in triplicate. solvent was transferred to a clean PP tube and mixed using a
The structure and morphology of the nanocomposite films vortex mixer (Scientific Industries Inc., NY, USA) for 30 s to
were characterized by X-ray diffraction (XRD) and trans- disperse the nanoclay particles. Approximately 2 mL of the
mission electronic microscopy (TEM). XRD analysis was solvent was transferred to a sample vial and an injection volume
carried out with a Bruker AXS D8 Advance X-ray diffractometer of 20 μL was used for instrumental analysis. After the analysis,
(Bruker Co., Billerica, MA, USA) equipped with a Global the sample solution was returned to the migration cell. All
Mirror filtered Cu Kα radiation source (wavelength, λ = 0.154 samples were tested in triplicate. Due to the small amount of
nm) setting at 40 kV and 40 mA. The film sample as well as nanoclay particles in the solvent, a “concentration function”
clay powder (control) was scanned over a 2 theta range of 0.5° associated with the instrument software was applied to ensure
to 10° at a rate of 0.5° min−1 and an increment of 0.01°. TEM better detection of the sample by increasing its signal. This
analysis was performed with a JEOL 100CX II TEM (JEOL function consists of repeating injections of the sample via the
USA Inc., MA, USA). The film sample was embedded in a auto sampler at the drying stage of the graphite furnace
paraffin block and cut with a microtome into 100 nm thin program. Each solvent sample was injected 6 times (20 μL per
sections. The microtomed sections were observed in a bright injection) before the data was recorded. The limit of
field imaging mode with an acceleration voltage of 120 kV. quantification (LOQ) was 8 μg L−1 for Si and 3 μg L−1 for Al.
Release Experiment for Polymer−Clay Films. Two- Electron Microscopy. The released nanoclay particles were
sided liquid extraction experiments were carried out in observed by bright field imaging using a JEOL JEM-2200FS
accordance with ASTM D4754-11. The apparatus used for field emission TEM (JEOL USA Inc., MA, USA) operating at
the experiment is shown in the Supporting Information (SI). an acceleration voltage of 200 kV. Composition analysis of the
Before the experiments, film samples were washed carefully observed particles was carried out with an X-ray energy
with water and ethanol to remove any contaminants on the dispersive spectroscopy (EDS) detector attached to the
polymer surface. Round disks (2 cm diameter) were cut from microscope. Once the release experiment was completed,
nanocomposite films (triplicate) as well as control films TEM analysis was carried out after obtaining 6 random samples
(triplicate). To test the release of nanoclay, a 50 mL PP tube of the solvent from the “migration cell” (PP tube) and pouring
was used as the migration cell; 16 film disks (total area = 100 them on a copper grid coated with a carbon membrane. The
cm2) of a single material were placed in the tube and the disks carbon membrane was used to collect and stabilize the particles,
were placed on the stainless steel wire and separated by Teflon and prevent the aggregation of the particles during drying and
beads. The tube was filled with 40 mL ethanol (200 proof or TEM observation. Three layers of filter paper were placed
100%) and kept at 70 °C for up to 10 d. Ethanol was used to below the copper grid to absorb the extra solvent. The copper
simulate fatty foods according to U.S. Food and Drug grid was dried under a 100W lamp for 10 min to evaporate the
Administration (FDA) recommendations on the migration residual solvent and observed under the microscope.
testing of food contact substances.31 A high temperature was Liquid Chromatography Tandem Mass Spectrometry.
used to accelerate the release process also according to the Quantification of surfactant in the solvent was performed by a
FDA recommendations.31 To test the release of surfactant, glass liquid chromatography tandem mass spectrometry (LC-MS/
vials (2.5 cm × 9.5 cm) were used instead of PP tubes as the MS) method as described in another study.33 A Waters Quattro
migration cell to avoid any absorption of surfactant by the tube; micro mass spectrometer (Waters Co., MA, USA) coupled to a
16 disks were placed in each vial as described above and the Shimadzu LC-20AD HPLC system (Shimadzu Scientific
extraction with ethanol was conducted at 70 °C for up to 12 d. Instruments, MO, USA) and a SIL 5000 autosampler were
For all experiments, multiple samplings were taken of the food used; the system was operated by using Waters MassLynx 4.0
simulant at varied time intervals until the end of the software. The instrumental conditions as well as the perform-
experiment. Nanoclay suspensions in ethanol made up of an ance of this method are briefly described in the Supporting
equivalent amount of nanoclay as in the nanocomposite films Information (SI).
were used as a control. Standard solutions of each surfactant in ethanol, with
Evaluation of Nanoclay Release. Nanoclay concentra- concentrations ranging from 0.1 to 5 mg L−1, were used to
tions in the solvent were determined by a graphite furnace establish the external calibration curve for each component of
atomic absorption spectrometry (GFAAS) method as described the surfactant. The LOQ was 25 μg L−1 for both surfactants
in our previous study.32 A Hitachi Z-9000 simultaneous when using C18C18 component as the marker. A sample aliquot
multielement atomic absorption spectrometer (Hitachi High- (1 mL) at each sampling time was transferred from the
Technologies Co., Tokyo, Japan) equipped with a HGA-700 migration cell to a 20 mL glass vial and diluted with 1 mL
atomizer and an autosampler system was used. Si and Al were ethanol for the Arquad surfactant or 2 mL ethanol for the
selected as markers for the nanoclays, and their contents in the Armeen surfactant. Meanwhile, at each sampling time the
nanoclays were evaluated via an X-ray fluorescence (XRF) migration cell was compensated with 1 mL ethanol. Each
method as also described in a previous study.32 XRF results for sample was filtered with a Waters GHP filter (13 mm, 0.2 μm;
the clays showed that I44P clay contained (20.14 ± 0.06) wt % Waters Co., MA, USA) before injection into the HPLC. Each
Si and (7.80 ± 0.02) wt % Al, with a Si/Al ratio of 2.58 ± 0.04; standard solution was injected 3 times, and each sample
whereas Cloisite clay contained (20.03 ± 0.16) wt % Si and solution was injected twice with an injection volume of 10 μL.
(7.15 ± 0.05) wt % Al, with a Si/Al ratio of 2.80 ± 0.12.
Calibration curves generated from Si and Al standards were
used to determine the amount of Si and Al released into the
■ RESULTS AND DISCUSSION
Properties of the Nanocomposite Films. For the
solvent. Nanoclay concentrations were calculated by correlating thermal properties, as shown in Table 1, no significant
to Si and Al concentrations based on the element contents in difference was found for Tg between the nanocomposite and
the nanoclay. control films, whereas Tm and Xc were significantly different.
13619 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624
Environmental Science & Technology Article

Table 1. Thermal Properties of the Nanocomposite and homogeneously dispersed in the PA6 matrix. A well exfoliated
Control Films structure can be achieved when there is good thermodynamic
affinity between the nanoclay and the polymer matrix;
films Tm (°C) Xc (%)a Tg (°C)
otherwise aggregation of nanoclay particles occurs if the
PP (control) 164.2 ± Ab
0.2 45.4 ± 1.0A 1.0 ± 1.2A polymer−clay interaction is thermodynamically unfavorable.38
PP−clay 162.7 ± 0.2B 42.3 ± 0.9B 1.4 ± 2.1A Therefore, better interaction was expected between Cloisite
PA6 (control) 220.2 ± 0.0A 24.8 ± 2.8A 61.5 ± 0.4A clay and PA6 than between I44P clay and PP.
PA6−clay 212.1 ± 0.2B 18.6 ± 2.0B 60.9 ± 0.2A Release of Nanoclay from Nanocomposite Films.
219.3 ± 0.3C Nanoclay released from nanocomposite films was determined
a
Heat of fusion of 100% crystalline PP36 is 207 J g−1 and heat of fusion by tracking the concentration of Si and Al in the solvent by
of 100% crystalline PA637 is 240 J g−1. bValues are the mean ± stdev; GFAAS as a function of time. The results are shown in Figure
student’s t test was applied to each property within each polymer type, 3. In order to be certain that the Si and Al measured originated
and means with different uppercase letters are significantly different (P from the nanoclay within the nanocomposite, two independent
< 0.05, n = 3).
evaluations were carried out. For the first evaluation, control
One extra Tm was obtained for PA6 at 212 °C after the addition films were run in parallel with the nanocomposite films. Si and
of nanoclay (DSC curves are shown in the Supporting Al concentrations in the solvent in contact with the control
Information); this peak referred to the γ-crystalline region, films were below the LOQ throughout the release experiment.
which is different from the α-crystalline region formed at 220 For the second evaluation, the Si/Al ratio was tracked as a
°C for pure PA6.34 The appearance of the lower Tm could be function of time throughout the release experiment, as shown
attributed to the good compatibility between the clay layers and in Figure 3. The Si/Al ratio was 2.62 ± 0.25 in the solvent in
PA6 matrix, which induces the formation of γ-crystal melting at contact with the PP−clay film and 2.84 ± 0.38 in the solvent in
lower temperatures.35 contact with the PA6−clay film. These values were in good
Nanoclays, when embedded into the polymer matrix, can be agreement with the ones obtained by XRF analysis (Si/Al ratio
intercalated or even exfoliated by the polymer chains depending of 2.58 ± 0.04 for I44P clay, and 2.80 ± 0.12 for Cloisite clay).
on the interaction between the nanoclay and the polymer. The Therefore, with these two assessments the contamination of Si
interaction can be determined by assessing the structure and and Al from other sources such as dust within the film did not
morphology of the nanocomposite. XRD patterns and TEM represent a concern and the main source of Si and Al was
images for both polymer−clay nanocomposites are shown in attributed to the nanoclay within the film.
Figure 2. The gallery distance (d-spacing) of I44P clay powder On the basis of the Si and Al concentrations in the solvent,
the release of nanoclay particles from the polymer−clay films
was confirmed. In addition, the increasing trend of Si and Al
concentrations at the beginning of the release experiment
indicated an initial release of nanoclay particles from the
nanocomposite films. Shortly thereafter a steady state was
reached, as there was no obvious increase in element
concentrations through the end of the experiment.
Al was considered a better marker of nanoclay than Si since
there were potentially more contamination sources for Si, such
as dusts or impurities in the nanoclay (e.g., quartz). Therefore,
Al concentrations were further converted to nanoclay
concentrations (Figure 4) on the basis of the original Al
content in each nanoclay (7.8% for I44P clay, and 7.2% for
Cloisite clay). Figure 4 shows that both PP−clay and PA6−clay
films released small amounts of nanoclay particles (approx-
imately 0.1−0.15 mg L−1). This result partially aligns with
Simon’s theory that large nanoparticles are difficult to release
from the polymer.39 However, in the current study, small but
significant amounts of clay were released. Additional assessment
Figure 2. XRD patterns for (a) PP−clay and (c) PA6−clay and TEM is needed to evaluate the size of the released particles. Although
images for (b) PP−clay and (d) PA6−clay. Scale bar: 200 nm. the total weight of nanoparticles released was small, the number
of particles and total surface area of such particles could be
(Figure 2a) was 2.66 nm (2θ = 3.32°) and increased to 3.19 nm large due to the nature of nanoparticles. Those particles may be
(2θ = 2.77°) after the nanoclay was embedded into PP. That less than 100 nm and in the size range of translocating and
increase was caused by the intercalation of polymer chains into penetrating cells.13,16 More nanoclay particles were released
the clay gallery. Partial aggregation of nanoclay particles was from PP−clay films than from PA6−clay films despite the fact
found in the PP matrix as shown by the small clusters in the that the initial nanoclay content in PP−clay films (3 wt %) was
corresponding TEM image (Figure 2b). The XRD patterns in less than that in PA6−clay films (5 wt %). Such a difference
Figure 2c for the PA6−clay nanocomposite and clay powder could be explained by the interaction between the nanoclay and
shows that nanoclay particles were well exfoliated since the the polymer. The exfoliation structure of Cloisite clay in PA6
peak observed for clay powder (2θ = 3.26°) disappeared in the had a larger surface area interacting with the polymer matrix.
nanocomposite pattern. The exfoliated structure was also The interaction was further mediated through the hydroxyl and
confirmed by the TEM image of the PA6−clay film (Figure 2d) amine groups where hydrogen bonding can form at the
in which the nanoclay particles are well separated and interface between the nanoclay and the polymer.9 Hence, there
13620 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624
Environmental Science & Technology Article

Figure 3. Amounts of Si and Al released from (a) PP−clay films and (b) PA6−clay films into ethanol as a function of time. The Si/Al ratio in the
solvent as a function of time is also shown. Fitted lines are included as a visual guide.

In contrast, poor interaction was found between I44P clay and


PP (although some PP was treated with MA to improve this
affinity) as indicated by the structure and morphology in Figure
2a & b. Therefore, a higher mobility was expected for I44P clay
particles, which increased their chance of release from the
polymer.
Characterization of Released Nanoclay Particles.
Figure 5 shows the nanoclay particles released into the solvent
as visualized by TEM. The dark areas within the circles in each
image consist of multiple parallel lines representing the stack of
clay layers (images with high magnification of the circled areas
are shown in the Supporting Information). EDS analysis
showed that the particle contains O (48.44 wt %), Si (33.07 wt
Figure 4. Concentration of nanoclay particles released from polymer− %), and Al (13.31 wt %), which are the major elements of
clay films into ethanol as a function of time. Nanoclay concentrations MMT nanoclay. The Si and Al contents, as determined by EDS
were further converted to milligrams of clay per squared meter of film analysis, were converted to the corresponding contents in the
and squared meters of clay per squared meter of film. The total surface I44P clay by multiplying by 65% (i.e., the MMT content in
area for exfoliated MMT nanoclay was assumed to be 750 m2 g−1.40 I44P clay; the remaining 35% is surfactant and does not contain
Fitted lines are included as a visual guide.
the three elements mentioned above). The resulting values
were 21.5% Si and 8.6% Al, with a Si/Al ratio of 2.5, similar to
would be a stronger interface between Cloisite clay and PA6, the values by XRF analysis of pure I44P clay (20% Si and 7.8%
which significantly reduced the mobility of nanoclay particles. Al, with a Si/Al ratio of 2.6). The nanoclay particles observed

Figure 5. TEM images of released nanoclay particles from the PP−clay films and the corresponding EDS analysis for the particle in image (a). The
structures (multiple parallel lines) shown within the circles exhibit a stack of clay layers. Scale bar: 50 nm for image a, 200 nm for images b and c.

13621 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624


Environmental Science & Technology Article

under TEM are much larger (500−1000 nm) in one dimension composite films as well as from nanoclay in the suspension as a
than those in the polymer matrix, probably due to the function of time was calculated with the calibration curve. The
aggregation of small particles in the solvent after release from total amount of surfactant was interpreted as a summation of
the film. the three components (C16C16, C16C18, and C18C18), and the
Change of d-Spacing after Solvent Exposure. Figure 6 results are shown in Figure 7. At the beginning of the release
shows the results of an initial experiment where solvent experiment the amount of surfactant in the solvent increased
immediately for both nanocomposite films and then reached a
steady state after 8 d for the PP−clay film and after 12 h for the
PA-clay film, with no further increase of surfactant release. The
more rapid rate of release of surfactant from the PA6−clay film
could be due to the swelling of PA6 in ethanol as reported for
other nylon films,42 while PP has better resistance to the
solvent. The solvent easily penetrated the PA6 polymer matrix,
swelled the nanoclay particles, and interacted with the
surfactant. Furthermore, when comparing the amount of
surfactant released from the film and from the nanoclay
suspension (control), more surfactant was released from the
PA6−clay film than from the control. It was assumed that the
amount of surfactant released from the nanocomposite film
would not be greater than that released from the corresponding
control, because of the probable absorption of surfactant by the
polymer. The unusual phenomenon for the PA6−clay film can
Figure 6. Change of d-spacing of nanoclay in a PP−clay film after be explained by the large interfacial forces between the
immersion in ethanol at 70 °C for 2 h and then exposed to air at room
nanoclay and the polymer as demonstrated previously
temperature for 0 h, 12 h, or 7 d. Control represents PP−clay film
before immersion in ethanol, and powder represents dry clay powder. (exfoliated structure of the nanocomposite and the formation
Values shown are for one replicate. of hydrogen bonding between the nanoclay and the polymer).
Such interfacial forces facilitate the exfoliation of the nanoclay
layers, which promotes strong friction among the nanoclay
exposure caused structural changes to the nanoclay within the layers and the polymer during film processing at high
nanocomposite film (XRD patterns are shown in the temperatures above the thermal degradation temperature of
Supporting Information). The d-spacing of the nanoclay in the O-MMT.43 The combination of friction and high
the PP−clay film increased from 3.19 to 3.42 nm after exposure temperature during film processing would promote the release
to ethanol at 70 °C for 2 h, due to the absorption of solvent by of extra surfactant from nanoclay surfaces into the polymer
nanoclay that expanded the clay gallery. After the film was matrix. Both nanocomposite films released much greater
removed from the solvent, the d-spacing decreased from 3.42 to amounts of surfactant (3.5 mg L−1 from PP−clay film and
2.75 nm over time, to a level that was even below the initial d- 16.2 mg L−1 from PA6−clay film) than nanoclay particles (0.15
spacing value (3.19 nm) before exposure to the solvent and mg L−1 from PP−clay film and 0.1 mg L−1 from PA6−clay film)
similar to the level of the clay powder. Excluding the effect of into the solvent. Nanoclay particles and surfactants differ in
solvent evaporation, the additional decrease of d-spacing was physical properties such as size and shape. The Connolly
assumed to be due to a dual effect of the release of surfactant solvent-excluded volume44 for both Arquad and Armeen
and the relaxation of polymer chains,9,41 which caused the surfactants (using C18C18 component) was calculated by
collapse of the clay gallery. Chem3D (ChemOffice Ultra 2002, CambridgeSoft Co., MA,
Release of Surfactant from Nanocomposite Films. The USA); the results show that the molecular volume of Arquad
amount of each surfactant component released from nano- surfactant is 667.1 Å3 or 0.667 nm3 (radius = 0.874 nm) and the

Figure 7. Total amount of surfactant released from (a) PP−clay films and control and (b) PA6−clay films and control into ethanol as a function of
time. Each control represents nanoclay suspended in ethanol, using an equivalent amount of nanoclay as in the corresponding nanocomposite films.
Fitted lines are included as a visual guide.

13622 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624


Environmental Science & Technology


Article

volume of Armeen surfactant is 655.0 Å3 or 0.655 nm3 (radius = REFERENCES


0.868 nm). The range of the surfactant volume is similar to the (1) Global Markets for Nanocomposites, Nanoparticles, Nanoclays, and
molecular volume of the surfactants reported by Zhao et al.45 of Nanotubes; BCC Research LLC: Wellesley, MA, USA, 2014; http://
around 0.6 nm3 (radius < 1 nm). These values are significantly www.bccresearch.com/ market -res earch/nanotechnology/
smaller than the size of nanoclay particles, from a few tens of nanocomposites-market-nan021f.html.
nm to hundreds of nm in the lateral dimension. Compared with (2) Han, K. Q.; Yu, M. H. Study of the preparation and properties of
nanoclay particles, the surfactant molecules can more easily fabrics of a PET/TiO2 nanocomposite prepared by in situ
move within the polymer, and their release may follow the polycondensation. J. Appl. Polym. Sci. 2006, 100, 1588−1593.
diffusion behavior of small molecules within the polymer matrix (3) Radheshkumar, C.; Munstedt, H. Antimicrobial polymers from
due to the presence of free volume and polymer chain polypropylene/silver composites-Ag+ release measured by anode
relaxation. stripping voltammetry. React. Funct. Polym. 2006, 66, 780−788.
The release of both nanoclay particles and surfactant from (4) Kashiwagi, T.; Grulke, E.; Hilding, J.; Groth, K.; Harris, R.;
Butler, K.; Shields, J.; Kharchenko, S.; Douglas, J. Thermal and
two nanocomposites was confirmed in this study. The release flammability properties of polypropylene/carbon nanotube nano-
process may occur during the manufacture, use and disposal of composites. Polymer 2004, 45, 4227−4239.
nanocomposites, thereby potentially exposing different environ- (5) Bal, S.; Samal, S. S. Carbon nanotube reinforced polymer
ments and biological systems to the nanocomposite compo- composites - A state of the art. Bull. Mater. Sci. 2007, 30, 379−386.
nents. The instrumental methodologies used herein for the (6) Pereira de Abreu, D. A.; Paseiro Losada, P.; Angulo, I.; Cruz, J. M.
measurement of nanoclay particles and surfactants in a food Development of new polyolefin films with nanoclays for application in
simulant can be expanded to evaluate release in environmental food packaging. Eur. Polym. J. 2007, 43, 2229−2243.
samples. In general, the release of both nanoclay particles and (7) Rathi, S.; Dahiya, J. B. Polyamide 66/nanoclay composite:
surfactants may present a safety concern. To the best of the synthesis, thermal and flammability properties. Adv. Mater. Lett. 2012,
authors’ knowledge, this is the first time that the entire release 3, 381−387.
profile of nanoclay and surfactant from nanocomposites is (8) Patel, H. A.; Somani, R. S.; Bajaj, H. C.; Jasra, R. V. Nanoclays for
polymer nanocomposites, paints, inks, greases and cosmetics
reported and correlated with interactions among the polymer,
formulations, drug delivery vehicle and waste water treatment. Bull.
the nanoclay, and the solvent. This information should be Mater. Sci. 2006, 29, 133−145.
useful to assess the risks associated with the exposure to (9) Sinha Ray, S.; Okamoto, M. Polymer/layered silicate nano-
nanoclay and surfactant, although further assumptions are composites: a review from preparation to processing. Prog. Polym. Sci.
needed to translate the experimental results to the actual 2003, 28, 1539−1641.
exposure dose. (10) De A. Prado, L. A. S.; Karthikeyan, C. S.; Schulte, K.; Nunes, S.


*
ASSOCIATED CONTENT
S Supporting Information
P.; De Torriani, I. L. Organic modification of layered silicates:
structural and thermal charaterizations. J. Non-Cryst. Solids 2005, 351,
970−975.
(11) Choudalakis, G.; Gotsis, A. D. Permeability of polymer/clay
Apparatus for the release experiment (Figure S1); brief nanocomposites: a review. Eur. Polym. J. 2009, 48, 967−984.
instrumental conditions for LC-MS/MS method for the (12) Lagaron, J. M.; Lopez-Rubio, A. Nanotechnology for bioplastics:
determination of the composition of the surfactants (Table opportunities, challenges and strategies. Trends Food Sci. Technol.
S1); extracted LC-MS/MS chromatographs of the three main 2011, 22, 611−617.
components of surfactants (Figure S2); DSC curves of PA6 and (13) Verma, N. K.; Moore, E.; Blau, W.; Volkov, Y.; Babu, P. R.
PA6−clay films (Figure S3); XRD patterns of PP−clay film Cytotoxicity evaluation of nanoclays in human epithelial cell line A549
after after solvent exposure (Figure S4); TEM images with using high content screening and real-time impedance analysis. J.
higher magnification showing the stacking of clay layers (Figure Nanopart. Res. 2012, 14, 1137−1147.
S5). This material is available free of charge via the Internet at (14) Baek, M.; Lee, J. A.; Choi, S. J. Toxicological effects of a cationic
http://pubs.acs.org. clay, montmorillonite in vitro and in vivo. Mol. Cell Toxicol. 2012, 8,


95−101.
(15) Lordan, S.; Kennedy, J. E.; Higginbotham, C. L. Cytotoxic
AUTHOR INFORMATION
effects induced by unmodified and organically modified nanoclays in
Corresponding Author the human hepatic HepG2 cell line. J. Appl. Toxicol. 2011, 31, 27−35.
*Phone: 517-353-0172. Fax: 517-353-8999. E-mail: mariar@ (16) Yamashita, K.; Yoshioka, Y.; Higashisaka, K.; et al. Silica and
msu.edu. titanium dioxide nanoparticles cause pregnancy complications in mice.
Notes Nat. Nanotechnol. 2011, 6, 321−328.
(17) Talmage, S. S. Environmental and human safety of major
The authors declare no competing financial interest.


surfactants: alcohol ethoxylates and alkylphenol ethoxylates. A report to
the Soap and Detergent Association; CRE Press: Boca Raton, FL, USA,
ACKNOWLEDGMENTS 1994.
This work was financially supported by the Center for (18) Venhus, S. H.; Mehrvar, M. Health effects, environmental
Packaging Innovation and Sustainability at the School of impacts, and photochemical degradation of selected surfactants in
Packaging, Michigan State University, and by the USDA water. Int. J. Photoenergy 2004, 6, 115−125.
National Institute of Food and Agriculture and Michigan (19) Ying, G. G. Fate, behavior and effects of surfactants and their
degradation products in the environment. Environ. Int. 2006, 32, 417−
AgBioResearch, Hatch projects M. Rubino and R. Auras. The
431.
support of the Agriculture and Food Research Initiative (20) Sonnenschein, C.; Soto, A. M. An updated review of
Competitive Grant no. 2014-67021-21630 from the USDA environmental estrogen and androgen mimics and antagonists. J.
NIFA provide funding for a conference presentation of this Steroid Biochem. Mol. Biol. 1998, 65, 143−150.
research. The authors would like to thank Lijun Chen from the (21) Routledge, E. J.; Sumpter, J. P. Estrogenic activity of surfactants
Department of Biochemistry at MSU for assistance with LC- and some of their degradation products assessed using a recombinant
MS/MS equipment. yeast screen. Environ. Toxicol. Chem. 2009, 15, 241−248.

13623 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624


Environmental Science & Technology Article

(22) Chaudhry, Q.; Scotter, M.; Blackburn, J.; Ross, B.; Boxall, A.; (42) McNally, T.; McNally, G. M.; Ahmad, M. N.; Murphy, W. R.;
Castle, L.; Aitken, R.; Watkins, R. Applications and implications of Kennedy, R. Sorption and diffusion of fuel components in various
nanotechnologies for the food sector. Food Addit. Contam. 2008, 25, nylons. In SPE/ANTEC’97 Conference Proceeding, Toronto, ON, Apr
241−258. 27−May 2, 1997; Volume III, pp 2794−2799.
(23) Silvestre, C.; Duraccio, D.; Cimmino, S. Food packaging based (43) Cervantes-Uc, J. M.; Cauich-Rodriguez, J. V.; Vazquez-Torres,
on polymer nanomaterials. Prog. Polym. Sci. 2011, 36, 1766−1782. H.; Garfias-Mesias, L. F.; Paul, D. R. Thermal degradation of
(24) Gottschalk, F.; Nowack, B. The release of engineered commercially available organoclays studied by TGA−FTIR. Thermo-
nanomaterials to the environment. J. Environ. Monit. 2011, 13, chim. Acta 2007, 457, 92−102.
1145−1155. (44) Connolly, M. L. Computation of molecular volume. J. Am.
(25) Raynor, P. C.; Cebula, J. I.; Spanqenberger, J. S.; Olson, B. A.; Chem. Soc. 1985, 107, 1118−1124.
Dasch, J. M.; D’Arcy, J. B. Assessing potential nanoparticles release (45) Zhao, F.; Du, Y.; Xu, J.; Liu, S. Determination of surfactant
during nanocomposite shredding using direct-reading instruments. J. molecular volume by atomic force microscopy. Colloid J. 2006, 68,
Occup. Environ. Hyg. 2012, 9, 1−13. 784−787.
(26) Thomas, T.; Thomas, K.; Sadrieh, N.; Savage, N.; Adair, P.;
Bronaugh, R. Research strategies for safety evaluation of nanomateri-
als, Part VII: Evaluating consumer exposures to nanoscale materials.
Toxicol. Sci. 2006, 91, 14−19.
(27) European Food Safety Authority (EFSA).. Guidance on the risk
assessment of the application of nanoscience and nanotechnologies in
the food and feed chain. EFSA J. 2011, 9, 2140 http://www.efsa.
europa.eu/en/efsajournal/doc/2140.pdf.
(28) Szakal, C.; Roberts, S. M.; Westerhoff, P.; Bartholomaeus, A.;
Buck, N.; Illuminato, I.; Canady, R.; Rogers, M. Measurement of
nanomaterials in foods: Integrative consideration of challenges and
future prospects. ACS Nano 2014, 8, 3128−3135.
(29) National Research Council (NRC). Research Progress on
Environmental, Health, and Safety Aspects of Engineered Nanomaterials;
The National Academies Press: Washington, DC, USA, 2013; http://
www.nap.edu/catalog.php?record_id=18475.
(30) Reichert, P.; Nitz, H.; Klinke, S.; Brandsch, R.; Thomann, R.;
Mülhaupt, R. Poly(propylene)/organoclay nanocomposite formation.
Macromol. Mater. Eng. 2000, 275, 8−17.
(31) Guidance for Industry: Preparation of Food Contact Notifications
and Food Additive Petitions for Food Contact Substances: Chemistry
Recommendations; The Food and Drug Administration (FDA): Silver
Spring, MD, USA, 2007; http://www.fda.gov/Food/
GuidanceRegulation/ucm081818.htm.
(32) Xia, Y.; Rubino, M.; Auras, R. Detection and quantification of
montmorillonite in water-ethanol solutions by graphite furnace atomic
absorption spectrometry. Food Addit. Contam. Part A 2013, 30, 2177−
2183.
(33) Xia, Y.; Rubino, M.; Auras, R. LC-MS/MS assay for the
determination of surfactants released from montmorillonite nanoclay
into food simulants. Appl. Clay Sci. 2014, submitted and revised.
(34) Katoh, Y.; Okamoto, M. Crystallization controlled by layered
silicates in nylon 6-clay nano-composite. Polymer 2009, 50, 4718−
4726.
(35) Wan, T.; Du, T.; Wang, B.; Zeng, W.; Clifford, M.
Microstructure, crystallization and dynamic mechanical properties of
polyamide/clay nanocomposites after melt-state annealing. Polym.
Compos. 2012, 12, 2271−2276.
(36) Vander Wal, A.; Mulder, J. J.; Gaymans, R. J. Facture of
polypropylene: 2. The effect of crystallinity. Polymer 1998, 39, 5477−
5481.
(37) Illers, K. H. Polymorphie, Kristallinität und Schmelzwärme von
Poly(ε-caprolactam), 2. Kalorimetrische Untersuchungen. Macromol.
Chem. 1978, 179, 497−507.
(38) Paul, D. R.; Robeson, L. M. Polymer nanotechnology:
Nanocomposites. Polymer 2008, 49, 3187−3204.
(39) Simon, P.; Chaudhry, Q.; Bakos, D. Migration of engineered
nanoparticles from polymer packaging to food − a physicochemical
view. J. Food Nutr. Res. 2008, 47, 105−113.
(40) Nikolaidis, A. K.; Achilias, D. S.; Karayannidis, G. P. Synthesis
and characterization of PMMA/organomodified montmorillonite
nanocomposites prepared by in situ bulk polymerization. Ind. Eng.
Chem. Res. 2011, 50, 571−579.
(41) Giannelis, E. P.; Krishnamoorti, R.; Manias, E. Polymer-silicate
nanocomposites: model systems for confined polymers and polymer
brushes. Adv. Polym. Sci. 1999, 138, 107−147.

13624 dx.doi.org/10.1021/es502622c | Environ. Sci. Technol. 2014, 48, 13617−13624

You might also like