You are on page 1of 9

European Polymer Journal 91 (2017) 212–220

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

ε-Caprolactone as a medium for improving dispersability of MARK


graphene oxide in polyamide based composites

Jaroslav Minář , Jiří Brožek
Department of Polymers, University of Chemistry and Technology Prague, Technická 5, 166 28 Prague 6, Czech Republic

AR TI CLE I NF O AB S T R A CT

Keywords: We described preparation of composites of graphene oxide with copolymers containing ε-


Polyesteramide caprolactam and ε-caprolactone structural units. Graphene oxide was prepared using modified
Graphene oxide Hummers method and characterized by elemental analysis and X-ray photoelectron spectroscopy
Composite measurements, interlayer distance was evaluated from X-ray diffraction. Two types of copoly-
In situ intercalation polymerization
mers - polyesteramides - with molar ratio of ε-caprolactam to ε-caprolactone units in
polymerization feed 80/20 and 90/10 were prepared by polymerization casting method.
Graphene oxide content in both types of polyesteramides was in range from 0.1 to 1 wt.%.
Thermogravimetric analysis and differential scanning calorimetry were utilized to study thermal
stability and crystallization behaviour of materials. Prepared composites were characterized by
tensile modulus and impact strength. Rheological properties of polymer melts and viscometry of
diluted polymer solutions revealed the physical character of polymer-filler interactions.

1. Introduction

Graphene oxide (GO) is layered material based on carbon, with history spanning over one and half century [1,2]. GO is composed
from graphene monolayers, constituted by sp2/sp3 carbon atoms arranged in a hexagonal lattice and containing oxygen functional
groups. Carbon to oxygen ratio depends on the preparation method and is roughly equal to 2/1 [3]. Although exact structure of GO is
still subject of discussion, many theoretical models assume presence of oxygen mostly in epoxide and tertial hydroxyl groups
covalently bonded to GO surface. Minor oxygen portion, however very important for chemical functionalization, is incorporated in
carboxyl groups located at monolayer peripheries [4]. Chemical, thermal or microwave reduction of GO leads to elimination of
significant amount of oxygen functional groups [5,6]; resulting material is denoted as “reduced graphene oxide”. GO and its
derivatives found various applications, e.g. in sensors [7], for “spin-coating” preparation of thin films [8,9], energy storage [10] and
many others [11].
Materials based on graphene are commonly used as a filler in polymer nanocomposites to improve properties of final product.
When high level of exfoliation and dispergation of layers is achieved, resultant materials are characterized by increased tensile
modulus, decreased gas permeability, increased thermal and electrical conductivity and other utility properties improvements [2].
Frequent method of nanocomposite preparation is in situ intercalation polymerization [12], where polymer chain growth is realized
in dispersion of filler in monomer melt. Covalent bonds between polymer and filler are often formed, which leads to superior
exfoliation and dispergation of filler in polymer matrix [2].
Poly(ε-caprolactam), also known as polyamide 6 (PA 6), belongs to the group of engineering plastics and is widely used for its
high tensile strength, high wear resistance and chemical resistance. PA 6 is prepared by polymerization of ε-caprolactam (CLA),


Corresponding author.
E-mail addresses: minarj@vscht.cz (J. Minář), Jiri.Brozek@vscht.cz (J. Brožek).

http://dx.doi.org/10.1016/j.eurpolymj.2017.03.058
Received 15 September 2016; Received in revised form 13 March 2017; Accepted 15 March 2017
Available online 30 March 2017
0014-3057/ © 2017 Elsevier Ltd. All rights reserved.
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

which can be carried out by two mechanisms - hydrolytic (polycondensation) and anionic [13]. The goal of preparation of PA 6 based
nanocomposites is to improve its properties. PA 6/GO nanocomposites were prepared in work [14] by hydrolytic mechanism.
Presence of only 0.1 wt.% in polymerization feed led to tensile modulus increase of 236%, grafting of polymer chains to GO surface
was confirmed by X-ray photoelectron spectroscopy and infrared spectroscopy with Fourier transformation. Anionic mechanism have
advantage over hydrolytic in high reaction rate and low content of low-molecular compounds in product. Therefore, this mechanism
is used in technology of polymerization casting or “reaction injection molding” (RIM) [15]. In situ intercalation via anionic
mechanism was applied in preparation of PA 6 nanocomposites containing up to 0.05 wt.% of modified GO [16]. Reaction was
initiated with sodium hydroxide, toluen-2,4-diisocyanate was used as an activator. Addition of GO caused increase in tensile strength
and Shore hardness of 111 and 116% respectively at the highest loading of filler. Simultaneously, coefficient of friction decreased by
10%.
Our laboratory tests demonstrated difficulties in dispergation of GO in CLA melt (as was also shown elsewhere [17]) leading to
inhomogenous distribution of filler in resulting materials and deterioration of their properties. Preliminary experiments proved that
addition of ε-caprolactone (CLO) into CLA/GO dispersion enhances dispersability of GO. Incorporation of CLO units into the PA 6
backbone leads to polyesteramides (PEAs) [18,19]. With increasing content of CLO units in PEAs, melting temperature, glass
transition temperature, Young’s modulus and tensile strength decreases. On the other hand, material becomes tougher.
The aim of this work is preparation and characterization of composites by in situ anionic copolymerization of CLA and CLO in
presence of GO.

2. Experimental

2.1. Materials

Graphite (Koh-i-noor Grafit s.r.o., Czech Republic). Sulfuric acid (96%), hydrogen peroxide (30%) and hydrochloric acid (35%)
(Penta s.r.o., Czech Republic). Potassium persulfate (99%) (Fluka). Phosphorus pentoxide (99,6%) (Lach-ner, Czech Republic).
Potassium permanganate (99,5%) (Lachema Praha, Czech Republic). ε-Caprolactam (Lanxess, Germany), water content determined
by Karl-Fischer titration was 150 ppm. ε-Caprolactone (Sigma-Aldrich) was dried over calcium hydride and distilled under the
reduced pressure, water content was 37 ppm. ε-Caprolactam magnesium bromide (CLAMgBr), concentrate in CLA (C1; c = 1.02 mol/
kg) was provided by Brügemann (Germany). Tricresol (Spolek pro chemickou a hutní výrobu, Czech Republic), mixture of o-, m- and
p-cresols was purified by four successive vacuum distillations and was stored in dark bottle under the argon atmosphere.

2.2. Preparation of graphene oxide

Two batches of GO (GO 1 and GO 2) were prepared by a modified Hummers method [1,14]. In the first step, graphite powder was
preoxidized by potassium persulfate and phosphorus pentoxide in sulfuric acid at 80 °C [14]. Treated graphite was subsequently
oxidized by pulverized potassium permanganate in sulfuric acid, temperature of reaction mixture did not exceed 10 °C. After slow
dosing of potassium permanganate, different temperature regimes were conducted in comparison to manual. Reaction mixture was
stirred for 2 h at laboratory temperature (GO 1) or at temperature maintained below 10 °C (GO 2). Afterwards, mixture was diluted
with 500 ml of deionized water (strong exothermic reaction indicated by sudden increase of temperature and evolving gases has
occured in both processes - in case of GO 1, this reaction took place after 20 min of stirring at laboratory temperature; in case of GO 2,
temperature raised after addition of about 50 ml of deionized water). After dilution, mixture was stirred at laboratory temperature
(both batches) for 2 h, following operations and separation of product were performed in accordance with manual [14].
Elemental analysis: GO 1 – C: 65.4%, H: 1.2%, calculated amount of oxygen: 33.4%; GO 2 - C: 60.6%, H: 1.6%, calculated amount
of oxygen: 37.8%.

2.3. Preparation of PEA/GO composites

Typical preparation of PEA composite with ratio of CLA/CLO in polymerization feed equal to 80/20, containing 0.1 wt.% of GO,
was depicted as follows. 0.06 g of GO 1 was dosed to 48 g of CLA melt. Flask containing CLA/GO dispersion under argon atmosphere
was placed in an ultrasonic bath and dispersion was sonicated for 2 h at 80 °C to reach higher level of dispergation of GO in monomer
melt. Dispersion was then heated to 110 °C and 3.786 g of polymerization initiator concentrate C1 (0.8 mol% of CLAMgBr in
polymerization feed) was loaded. After 30 min, which is the necessary time for complete dissolution of initiator, 12 g of CLO was
introduced into dispersion and followed by vigorous stirring, mixture was rapidly transferred into aluminium polymerization mould
(internal dimensions – 160 × 40 × 10 mm), preheated to 125 °C. Aluminium mould was immersed in oil bath heated to 150 °C,
polymerization was carried out for 1 h. According to that procedure, composites with molar ratio of CLA/CLO units 80/20 (notation -
PEA(80)) and 90/10 (notation - PEA(90)) were prepared. For preparation of PEA(80) composites, GO 1 was used. On the other hand,
GO 2 was used for PEA(90) composites. Contents of GO in both types of PEA were 0; 0.1; 0.6 and 1 wt.%.

2.4. Characterization

Elemental analysis of GO was performed on Elementar Vario EL III device (Elementar). Measurement accuracy was smaller than
0.1%.

213
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

Chemical composition of surface layers of GO was determined via X-ray photoelectron spectroscopy (XPS) on ESCAProbeP
(Omicron Nanotechnology Ltd.). As a source of monochromatic light, Al anode with energy 1486.7 eV was applied.
Interlayer distance of GO layers were measured by X-ray diffraction (XRD) on X’Pert3 Powder diffractometer (PANanalytical)
equipped with a Cu Kα tube providing wavelength of λ = 0,15406 nm.
Polymer yield was determined gravimetrically after extraction of low-molecular compounds, performed three times for 20 min in
boiling distilled water (1 g of sample per 100 ml).
Differential scanning calorimetry (DSC) measurements were carried out on DSC Q100 device (TA Instruments). Scanning rate was
10 °C/min, measurement was performed in the temperature range 0–220 °C in mode heating–cooling. Nitrogen was used as a sample
gas with a flow of 50 ml/min.
Thermal stability was evaluated from thermogravimetric analysis (TGA) measurement performed on instrument TGA Q500 (TA
Instruments) under nitrogen atmosphere (gas flow 60 ml/min) at a heating rate of 20 °C/min in temperature range from 20 to 600 °C.
For Transmission electron microscopy (TEM) measurements, ultrathin sections of nanocomposites were obtained using Leica EM
UC 6 at room temperature. TEM measurements were carried out on EFTEM Jeol 2200 FS at an acceleration of 200 kV.
Viscosimetric measurements were carried out at 25 °C in Übbelohde viscosimeter with a control unit VistecTM. Diluted solutions of
PEA composites in tricresol (approximate concentration 2 × 10−3 g/ml) were heated at 50 °C for 30 min to disrupt polymer
associates. In order to remove solid and gel-like particles, solutions were transferred into the viscosimeter through sintered glass filter
(porosity 90–160 μm). Limiting viscosity number [η] was calculated from equation [η] = [(1 + 1.6⋅ηi)0.5 − 1]/(0.8⋅c), Ref. [20],
valid for PA 6/tricresol system, where ηi represents increment of relative viscosity and c represents concentration of solution (g/ml).
Calculations of apparent viscosity average molar masses Mv were based on equation Mv = 113.1⋅118.5⋅ (0.01⋅[η])1.35, Ref. [21].
Standard deviation of the measurement did not exceed 6%.
Dynamic viscosities of composite melts were investigated on AR-G2 device (TA Instruments), using plate-plate arrangement.
Plates diameter was 25 mm, distance between plates was 0.7 mm. Measurements were carried out from the temperature 20 °C above
the melting temperature up to 300 °C at a heating rate of 5 °C/min. Shear deformation rate was adjusted to 0.1 s−1.
For determination of mechanical properties, test specimens were prepared from casted slabs. Dimensions of specimens were
140 × 4 × 2 mm (Young’s modulus) and 50 × 6 × 4 mm (impact strength). Specimens were stored in desiccator over the
phosphorus pentoxide.
Tensile tests were performed on Instron 3365 device at a rate of 1 mm/min until the proportional elongation exceeded 1%. Impact
strength was measured on Ceast Resil 5.5 instrument according to standard method ISO 179.

3. Results and discussion

3.1. Preparation of graphene oxide

Chemical composition of GOs prepared by modified Hummers method was studied with XPS. Fig. 1 shows full-range XPS spectra
with distinctive C1s peak at 285.1 eV (GO 1, GO 2) and O1s peak at 533.1 (GO 1) or 532.7 (GO 2) eV. Deconvoluted C1s component
peaks are shown in Fig. 2. Their positions and areas, that corresponds to the mass portions of related functional groups, are
summarized in Table 1.
Elemental analysis gives the C/O ratio equal to 1.96 (GO 1) and 1.60 (GO 2). However, peak area ratios C1s/O1s based on XPS are
3.47 and 3.49. Shape of component peaks (Fig. 2) also fits better to lower oxidation degree GOs [22]. These variations originate from
different character of mentioned analysis. Elemental analysis measures total composition of entire sample, whereas XPS is surface
method, that is able to describe only few layers.
Interlayer distance of GO layers was measured by XRD. It is known that introduction of oxygen functional groups to graphite
structure increases its interlayer distance from 0.34 nm to values about 0.7–0.8 nm [23]. Both types of GO exhibited interlayer
distance of only 0.38 nm. Furthermore, GO 2 had a diffraction peak corresponding to 0.34 nm. Moderate increase of interlayer
distance can be attributed to low content of epoxide functional groups (see Table 1), which are responsible for increasing interlayer

Fig. 1. XPS spectra of GO 1 (a) and GO 2 (b).

214
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

Fig. 2. Deconvoluted C1s peaks from XPS spectra of GO 1 (a) and GO 2 (b).

distance of GO [22]. Small differences of prepared GOs are not expected to have a significant impact on properties of resulting
composites.

3.2. Preparation of composites

Composites were prepared by anionic copolymerization of CLA and CLO mixture that contains dispergated GO. Advantage of
anionic polymerization is its high reaction rate. On the other hand, this reaction is very sensitive to the presence of polar impurities,
because transfer of hydrogen proton can terminate polymer chain growth [24]. GO 1 and GO 2 contains mostly carboxyl and
hydroxyl functional groups and bounded water, which can negatively affect polymerization process. According to literature [25], it is
known, that polymerization initiator CLAMgBr is less sensitive to the presence of polar impurities, compared to conventional initiator
sodium ε-caprolactamate. CLAMgBr undergoes disproportionation reaction, yielding magnesium di(ε-caprolactamate) (CLA2Mg) and
magnesium bromide, see Eq. (1). Subsequent reaction of CLA2Mg with water results in formation of ε-caprolactam magnesium
hydroxides CLAMg(OH) (Eq. (2)). Their condensation to magnesium hydroxide (Eq. (3)) decreases water content in polymerization
system. CLA2Mg can also react with carboxyl and hydroxyl functional groups of GO, causing formation of magnesium carboxylates
and hydroxylates (Eq. (4)). These reactions also causes a decrease of water content in polymerization feed. Therefore, increasing
concentration of GO in polymerization feed had to be accompanied by increasing concentration of CLAMgBr, see Tables 2 and 3.
Initiator concentrations were adjusted on basis of preliminary polymerization tests with objective of high degree of monomer
conversion.

Composites were prepared via polymerization casting method, see Section 2.3. Viscosity of CLA/GO dispersion containing
dissolved CLAMgBr sharply increased after addition of CLO. We assume, by analogy with literature [18], that favored reaction that
takes place after dosing of CLO, is its polymerization to poly(ε-caprolactone), even during the preparation of polymerizaton feed. The
participation of CLA units in the polymerization reaction occurs at higher temperatures. Transacylation reactions takes place
throughout the ongoing polymerization, resulting in random character of copolymer. Another benefit of anionic over hydrolytic
polymerization is lower reaction temperature, which is 150 °C (hydrolytic polymerization is usually carried out at 260 °C), when the

Table 1
Positions and areas of deconvoluted C1s peaks of GO 1 and GO 2.

Functional group GO 1 position (eV) GO 2 position (eV) GO 1 peak area (%) GO 2 peak area (%)

eCH2e 284.8 284.7 65.4 71.2


eCeOe 286.1 286.6 16.9 16.6
eC]O (eOeCeO) 287.1 287.8 2.9 1.2
eOeC]O 288.8 288.9 14.8 11.0

215
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

Table 2
Composition of polymerization feed and properties of PEA(80) composites prepared by polymerization casting.

Sample No. cGOa (wt.%) cCLAMgBrb (mol%) ywc (%) Etd (MPa) ake (kJ/m2) ΔHm/Tmf/g (J/g)/(°C)

PEA(80)-0 0 0.80 96.0 1186 ± 91 14.41 ± 0.57 40/177


PEA(80)-0.1 0.1 0.80 96.2 1109 ± 35 13.98 ± 0.87 42/176
PEA(80)-0.6 0.6 1.60 95.1 1463 ± 67 3.22 ± 0.25 46/176
PEA(80)-1 1 2.00 94.9 1798 ± 72 2.92 ± 0.24 44/178

a
cGO – concentration of GO in polymerization feed.
b
cCLAMgBr – concentration of CLAMgBr in polymerization feed.
c
yw – polymer yield.
d
Et – Young’s modulus.
e
ak – impact strength.
f
ΔHm – enthalpy of fusion.
g
Tm – melting temperature.

mass loss of GO is insignificant, see Fig. 3. Thanks to this feature, casted slabs did not contain any cavities or surface defects, arising
from decomposition of GO functional groups. High degree of monomer conversion (> 94%) was observed, see Tables 2 and 3.
Therefore, we could prepare test specimens for subsequent analysis directly from slabs, without necessity of removing low-molecular
compounds.

3.3. Thermal behaviour of GO and composites

TGA records of GO and PEA composites are shown in Fig. 3. GO is thermally unstable compound and gradual mass loss occurs due
to the liberation of bounded water (up to 150 °C) and decomposition of oxygen functional groups in whole temperature range of
analysis [5]. Shift of GO decomposition curves is related to carbon content and is in accordance with elemental analysis results, see
Section 2.2.
Effect of graphene-based materials on decomposition behaviour of polymer nanocomposites is an ambiguous problem. Several
studies have found that presence of graphene oxide enhances thermal stability of nanocomposites [14,26,27]. In our case, however,
presence of GO is not beneficial for thermal stabilities of nanocomposites. Temperatures at the highest decomposition rate THDR (i.e.
maximum of the derivative curve) were selected as a comparative criterion of thermal stabilities, see Table 4. Addition of GO to PEA
matrix decreases THDR almost linearly, this effect being more noticeable for PEA(90) nanocomposites. An obvious contradiction to
literature might have several explanations. The first one lies in high oxygen content of GO’s, that reaches almost 40 wt.%. As a result,
pristine GO weight loss is around 15/20 % (GO 1/GO 2) at the PEA initial decomposition temperature. A second and more important
reason is that nanocomposites with higher content of GO were prepared using higher concentration of magnesium initiator (see
Section 3.2), therefore they contain more magnesium residues. High oxygen content of GO together with magnesium residues in
polymer matrix can deteriorate thermal stability of PEA nanocomposites.
Melting temperature, enthalpy of fusion and rate of crystallization were evaluated from DSC records, see Fig. 4. Presence of GO is
not crucial for values of enthalpy of fusion or melting temperatures, see Tables 2 and 3. Single endothermic peak of melting proves
statistic distribution of structural units along the polymer chain [18]. Fillers based on carbon act as a heterogenous nucleation agent
[2], resulting in shift of crystallization exotherm peaks of PEA composites to higher temperatures and decreasing their full width at
half maximum.

3.4. Transmission electron microscopy

Morphological observations of sample PEA(90)-0.6 were carried out using TEM, see Fig. 5. From Fig. 5(a) is obvious that GO is
homogenously dispersed throughout the PEA matrix. Low dispersion density might be attributed to low concentration of GO. Bright
region (highlighted by arrow) corresponds to the polymer crystalline phase. Its encapsulation by filler confirms role of GO as a
nucleation agent, see Section 3.3. More detailed look at the interactions between polymer and filler is proposed in Fig. 5(b).
Aggregated structure, which occurs near to the edge of ultramicrotome cut, suggests that GO layers are not fully exfoliated.

Table 3
Composition of polymerization feed and properties of PEA(90) composites prepared by polymerization casting.

Sample No. cGOa (wt.%) cCLAMgBrb (mol%) ywc (%) Etd (MPa) ake (kJ/m2) ΔHm/Tmf/g (J/g)/(°C)

PEA(90)-0 0 0.90 97.7 2209 ± 113 4.7 ± 0.2 60/199


PEA(90)-0.1 0.1 0.90 97.6 2253 ± 109 3.8 ± 0.1 62/198
PEA(90)-0.6 0.6 2.25 96.0 2394 ± 149 2.3 ± 0.3 62/198
PEA(90)-1 1 3.15 94.4 2539 ± 205 2 ± 0.3 62/199

Symbols see Table 2.

216
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

100

80

GO 2

Weight (%)
60
GO 1
PEA(90)-1
PEA(90)-0,6
PEA(90)-0
40 PEA(80)-1
PEA(80)-0,6
PEA(80)-0

20

0
0 200 400 600
Temperature (°C)

Fig. 3. TGA curves of GO 1 and GO 2, PEA(80) and PEA(90) composites with GO content of 0, 0.6 and 1 wt.% in polymerization feed.

3.5. Viscosimetric measurements

Apparent viscosity average molar masses (Mv) of PEA samples were determined using viscosimetric measurements, see Section
2.4. Samples were dissolved in tricresol and prior to the measurements were filtered through sintered glass filter. Gelled particles,
giving proof of presence of covalent bonds between polymer chains and filler or formation of cross-linked structures, were not
detected on filter. PEA(90)-0 exhibits higher Mv compared to PEA(80)-0, due to the lower content of CLO units, see Fig. 6(a). Presence
of only 0.6 wt.% of GO in both PEA types decreases Mv to values around 16,000 g/mol and it almost does not change at concentration
of 1 wt.% of GO. This decrease can be explained by hydrolysis of ester bonds by water present in GO or water released by
condensation of hydroxyl groups at polymerization temperature. Magnesium compounds cannot effectively absorb all water arising
from GO as we proposed in Section 3.2.
Dependences of viscosities of polymer melts on temperature are shown in Fig. 6(b). Melt viscosity is influenced by mutual
interactions of polymer chains, or more precisely of their functional groups, and by molar mass of polymer. Lower Mv of PEA(80)-0 in
comparison with PEA(90)-0 corresponds to shift of viscosity curves. Drop of viscosity curves of PEAs in presence of GO is in
agreement with Mv. Viscosity increase at temperatures higher than 270 °C is connected with degradation of PEA chains caused by
raised temperature.
If carboxyl groups of GO were reacting with PEA ester groups, so covalent bonds were created. Resulting particles would have
bigger hydrodynamic volume and as a result, molar mass would be increasing. Viscosimetric measurements then implies, that filler is
not covalently bonded to polymer chains.

3.6. Mechanical properties of composites

When good level of dispergation (intercalation or exfoliation) of layered fillers in polymer matrix is achieved, we can expect
positive influence on many mechanical properties in contrary to unmodified materials [2,14,28]. In our experiment, presence of
0.1 wt.% in polymerization feed does not affect significantly Young’s modulus of PEA composites, see Tables 2 and 3 and Fig. 7(a).
With increasing GO concentration, we can observe increase in Young’s modulus of 52% in case of PEA(80)-1. In case of PEA(90)-1,
modulus increase is only 15%.
Similarly to Young’s modulus, 0.1 wt.% of GO does not influence impact strength of PEA composites (Fig. 7(b), Tables 2 and 3).
Impact strength of both types of composites then strongly decreases with increasing GO concentration on values around 2–3 kJ/m2.
Viscosimetric measurements indicate the lack of covalent bonds between PEA chains and GO functional groups. Therefore, we
have to explain influence of GO on mechanical properties of composites by different kind of polymer-filler interactions. We suppose
contribution of hydrogen bonds between hydroxyl or carboxyl groups of GO and PEA amidic groups, as it is suggested in work [29].
Magnesium compounds emerging from reactions of initiator in CLA/GO dispersion (see Eqs. (1)-(3)) play crucial role in explaining
this phenomenon. Their reaction with functional groups of GO constitutes magnesium carboxylates or hydroxylates (see Eq. (4)). It is
generally known, that magnesium ions are creating complex compounds with CLA molecules [30] or with amide groups of polyamide
chains through its carbonyl groups [31]. This is why we suggest that magnesium ions bounded to GO are creating complex structures

Table 4
Temperatures at the highest decomposition rate (THDR) of PEA(80) and PEA(90) nanocomposites.

GO content 0 0.6 1

THDR of PEA(80) (°C) 405 401 396


THDR of PEA(90) (°C) 409 401 388

217
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

0.0 1.8

1.6
-0.2 PEA(90)-1
1.4
PEA(90)-0,6
PEA(90)-0
PEA(80)-1
-0.4 1.2 PEA(80)-0,6
PEA(80)-0

Heat Flow (W/g)


Heat Flow (W/g)

1.0
-0.6 PEA(90)-1
PEA(90)-0,6 0.8
PEA(90)-0
PEA(80)-1
-0.8 PEA(80)-0,6 0.6
PEA(80)-0
0.4
-1.0
0.2

-1.2 0.0
100 150 200 80 130 180
Exo Up
Exo Up
Temperature (°C) Temperature (°C)

(a) (b)
Fig. 4. DSC records of PEA(80) and PEA(90) composites with GO content of 0, 0.6 and 1 wt.% in polymerization feed in heating (a) and cooling (b) mode.

Fig. 5. TEM micrographs of PEA(90)-0.6 sample.

with PEA chains, as is shown in Fig. 8.


Those interactions are responsible for increase in Young’s modulus of composites. Interactions can appear only at polymer chains
located in an amorphous state. PEA(80) composites includes higher portion of amorphous phase due to higher content of CLO units,
see Tables 2 and 3. Therefore, the modulus increase of PEA(80) composites is steeper as a consequence of higher quantity of
interactions, compared to PEA(90) composites.

4. Conclusions

GO was prepared by modified Hummers method. Transformation of graphite to GO became evident in increase of oxygen content
in range of 30–40 %, as was confirmed by elemental analysis and XPS measurements.
Composites of PEA(80) and PEA(90) with GO content from 0.1 to 1 wt.% were prepared by polymerization casting method.
Gravimetric extraction confirmed low content of low-molecular portions.
Presence of only one melting temperature of PEAs proves statistical arrangement of CLA and CLO units in polymer matrix. GO acts
as a nucleating agent, which was deduced from accelerated rate of crystallization and increased crystallization temperature in the
mode of cooling of polymer melt during the DSC measurement.
GO has stiffening effect on polymer matrix, as follows from increase of Young’s modulus with increasing concentration of GO. On
the other hand, composites became more brittle with higher filler concentration, which was proved by significant decrease of impact

218
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

Fig. 6. (a) Dependence of apparent viscosity average molar mass (Mv) of PEA composites on content of GO. (b) Dependence of melt viscosity of PEA composites on
temperature.

Fig. 7. PEA(80) and PEA(90) composites dependences of Young’s modulus (a) and impact strength (b) on GO content in polymerization feed.

Fig. 8. Possible interactions of magnesium ions with carboxyl and hydroxyl group of GO and carbonyl groups of PEA chains, figure also shows hydrogen bond between
PEA chains.

strength.
Decrease both in melt viscosities and Mv of PEAs containing GO in contrary to unmodified ones gives a proof of physical nature of
polymer-filler interactions.

219
J. Minář, J. Brožek European Polymer Journal 91 (2017) 212–220

Acknowledgments

This work was supported from specific university research (MSMT No. 20-SVV/2016).

References

[1] W.S. Hummers Jr., R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80 (6) (1958) 1339.
[2] J.R. Potts, et al., Graphene-based polymer nanocomposites, Polymer 52 (1) (2011) 5–25.
[3] H. Kim, A.A. Abdala, C.W. Macosko, Graphene/polymer nanocomposites, Macromolecules 43 (16) (2010) 6515–6530.
[4] A. Lerf, et al., Structure of graphite oxide revisited, J. Phys. Chem. B 102 (23) (1998) 4477–4482.
[5] C.H.A. Wong, et al., Vacuum-assisted microwave reduction/exfoliation of graphite oxide and the influence of precursor graphite oxide, Carbon 77 (2014)
508–517.
[6] D.R. Dreyer, et al., The chemistry of graphene oxide, Chem. Soc. Rev. 39 (1) (2010) 228–240.
[7] J.T. Robinson, et al., Reduced graphene oxide molecular sensors, Nano Lett. 8 (10) (2008) 3137–3140.
[8] H.A. Becerril, et al., Evaluation of solution-processed reduced graphene oxide films as transparent conductors, ACS Nano 2 (3) (2008) 463–470.
[9] X. Wang, L. Zhi, K. Müllen, Transparent, conductive graphene electrodes for dye-sensitized solar cells, Nano Lett. 8 (1) (2008) 323–327.
[10] B. Luo, et al., Two dimensional graphene-SnS2 hybrids with superior rate capability for lithium ion storage, Energy Environ. Sci. 5 (1) (2012) 5226–5230.
[11] Y. Zhu, et al., Graphene and graphene oxide: synthesis, properties, and applications, Adv. Mater. 22 (35) (2010) 3906–3924.
[12] E. Duguet, S. Rey, J.M.M. Robles, Intercalation polymerization, Encyclopedia of Polymer Science and Technology, John Wiley & Sons, Inc., 2003.
[13] R. Puffr, V. Kubánek, Lactam-based Polyamides, Volume I: Polymerization Structure, vol. 1, Boca Raton, CRC Press, 1991.
[14] Z. Xu, C. Gao, In situ polymerization approach to graphene-reinforced Nylon-6 composites, Macromolecules 43 (16) (2010) 6716–6723.
[15] R.M. Hedrick, J.D. Gabbert, M.H. Wohl, NYLON 6 RIM, in: ACS Symposium Series, 1985.
[16] B. Pan, et al., Tribological and mechanical investigation of MC nylon reinforced by modified graphene oxide, Wear 294–295 (2012) 395–401.
[17] P. Ding, et al., Highly thermal conductive composites with polyamide-6 covalently-grafted graphene by an in situ polymerization and thermal reduction process,
Carbon 66 (2014) 576–584.
[18] D. Chromcová, et al., Polymerization of lactams. 99 Preparation of polyesteramides by the anionic copolymerization of ε-caprolactam and ε-caprolactone, Eur.
Polym. J. 44 (6) (2008) 1733–1742.
[19] I. Goodman, R.N. Vachon, Copolyesteramides-II. Anionic copolymers of ε-caprolactam with ε-caprolactone. Preparation and general properties, Eur. Polym. J. 20
(6) (1984) 529–537.
[20] F. Rybníkář, Chem. Listy 49 (1965) 1142.
[21] C.V. Goebel, et al., Anionic polymerization of caprolactam. XLIII. Relationship between osmometric molecular weight, viscosity, and endgroups of a polymer, J.
Polym. Sci. Part A-1 Polym. Chem. 10 (5) (1972) 1411–1427.
[22] K. Krishnamoorthy, et al., The chemical and structural analysis of graphene oxide with different degrees of oxidation, Carbon 53 (2013) 38–49.
[23] M.J. McAllister, et al., Single sheet functionalized graphene by oxidation and thermal expansion of graphite, Chem. Mater. 19 (18) (2007) 4396–4404.
[24] G. Odian, Ionic chain polymerization, Principles of Polymerization, fourth ed., John Wiley & Sons, Inc., 2004.
[25] P. Bernat, et al., Polymerization of lactams. 98{star, open}Part 97: cf. Ref. [10].{star, open}: influence of water on the non-activated polymerization of ε-
caprolactam, Eur. Polym. J. 44 (1) (2008) 32–41.
[26] X. Li, et al., Enhanced thermal-conductive and anti-dripping properties of polyamide composites by 3D graphene structures at low filler content, Compos. Part A
Appl. Sci. Manuf. 88 (2016) 305–314.
[27] Y. Liu, Z. Chen, G. Yang, Synthesis and characterization of polyamide-6/graphite oxide nanocomposites, J. Mater. Sci. 46 (4) (2011) 882–888.
[28] P. Steurer, et al., Functionalized graphenes and thermoplastic nanocomposites based upon expanded graphite oxide, Macromol. Rapid Commun. 30 (4–5) (2009)
316–327.
[29] H.R. Pant, et al., Photocatalytic TiO2–RGO/nylon-6 spider-wave-like nano-nets via electrospinning and hydrothermal treatment, J. Membr. Sci. 429 (2013)
225–234.
[30] J. Kříž, et al., Molecular structure of the complex of hexano-6-lactam with magnesium bromide, Macromol. Chem. Phys. 202 (7) (2001) 1194–1199.
[31] P. Arnoldová, et al., Role of magnesium complexes in the anionic polymerization of hexano-6-lactam, E-Polymers (2006) 1–11.

220

You might also like