Sei sulla pagina 1di 52

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/282184871

Synthesis, Properties, and Applications of


Polymeric Nanocomposites , Sevan P. Davtyan,
Aleksandr Berlin, Vladimir Agabekov...

Article in Journal of Nanomaterials · August 2012


DOI: 10.1155/2012/215094

CITATION READS

1 31

4 authors, including:

Sevan Paruyr Davtyan Alexander Alexandrovich Berlin


State Engineering University of Armenia Russian Academy of Sciences
732 PUBLICATIONS 727 CITATIONS 563 PUBLICATIONS 1,222 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Biodegradable polymeric compositions View project

Gradient Hydrojels by frontal polymerization View project

All content following this page was uploaded by Sevan Paruyr Davtyan on 21 February 2015.

The user has requested enhancement of the downloaded file.


ISSN 20799780, Review Journal of Chemistry, 2013, Vol. 3, No. 1, pp. 1–51. © Pleiades Publishing, Ltd., 2013.
Original Russian Text © S.P. Davtyan, A.S. Avetisyan, A.A. Berlin, A.O. Tonoyan, 2013, published in Obzornyi Zhurnal po Khimii, 2013, Vol. 3, No. 1, pp. 000–000.

Synthesis and Properties of ParticleFilled


and Intercalated Polymer Nanocomposites
S. P. Davtyana, A. S. Avetisyana, A. A. Berlinb, and A. O. Tonoyana
a
State Engineering University of Armenia, ul. Teryana 105, Yerevan, 0009 Armenia
b
Semenov Institute of Chemical Physics, Russian Academy of Sciences, ul. Kosygina 4, Moscow, #119991 Russia
email: davtyans@seua.am, atonoyan@mail.ru
Received Aril 4, 2012; in final form, October 2, 2012

Abstract—This paper analyzes available data on the synthesis and properties of polymer nanocompos
ites prepared by various techniques (sol–gel processing and microemulsion and frontal polymeriza
tions) and containing polymethyl methacrylate, polydimethylsiloxane, natural rubber, and other poly
mers as binders, and various amounts of nano and microadditives: SiO2, TiO2, FeO, clay, and
Y1Ba2Cu3O7 – x. We consider the physicomechanical, dynamic mechanical, superconducting, ther
mophysical, thermochemical, and other properties of the polymer nanocomposites.
Also examined are data on the activated anionic polymerization of εcaprolactam in the presence of
various amounts of SiO2 nanoparticles. Results on crystallization kinetics and electron microscopy
data lead us to conclude that SiO2 nanoparticles act as heterogeneous nucleation centers for the crys
tallization of the forming poly(εcaprolactam).
Analysis of our results and data reported by other groups demonstrates that the intercalation of poly
mer macromolecules and their fragments into the interlayer spaces of ceramic grains increases the
superconducting transition temperature of the ceramic by 1–3°C.
Key words: intercalated polymer nanocomposites, rigid amorphous fraction, surface layer, frontal and
adiabatic polymerization, crystallization kinetics
DOI: 10.1134/S2079978013010019

CONTENTS
1. Introduction
2. ParticleFilled Polymer Nanocomposites
2.1. PMMA/SiO2, PMMA/Clay, PS/SiO2, and PS/Clay Nanocomposites Produced by Different Tech
niques and Their Characteristics
2.2. Microemulsion Polymerization
2.3. Formation of the Immobilized Polymer Fraction
2.4. PAA Nanocomposites Prepared under Different Thermal Conditions
2.5. Formation of the Rigid Amorphous Fraction under Frontal Polymerization Conditions and Thermo
physical Characteristics of the Resultant Nanocomposites
2.6. Dynamic Mechanical Properties of PMMA/SiO2 Nanocomposites
2.7. Effect of Nanoparticulate Additives on the Kinetics and Mechanism of the Radical Polymerization of
Vinyl Monomers
2.8. Effect of SiO2 Nanofiller on Crystallization Kinetics under Adiabatic εCaprolactam Anionic Polymer
ization Conditions
2.9. Thermodynamic Aspects of Nanoparticle–Macromolecule Interaction
3. Superconducting Intercalated Polymer–Ceramic Nanocomposites
3.1. Superconducting Properties of Polymer–Ceramic Nanocomposites Prepared by Hot Pressing
3.2. Effects of Thermal Conditions and Antioxidant Additives on ThermoOxidative Destruction Processes
3.3. Effect of ThermoOxidative Destruction on the Superconducting Properties of Polymer–Ceramic
Nanocomposites
3.4. Methyl Methacrylate Polymerization in the Presence of Ceramic Y1Ba2Cu3O7 – x
3.5. Fabrication of Superconducting Polymer–Ceramic Nanocomposites under the Conditions of the Fron
tal Polymerization of Mn, Co, Zn, and NiContaining Metal Complex Monomers in the Presence of
Y1Ba2Cu3O6.97

1
2 DAVTYAN et al.

3.6. Physicomechanical Properties of Superconducting Polymer–Ceramic Nanocomposites


3.7. Effects of Particle Size Composition and Filling Fraction on the Physicomechanical and Superconduct
ing Properties of Polymer–Ceramic Nanocomposites
3.8. Interfacial Phenomena in HighTc Polymer–Ceramic Nanocomposites
3.9. Dynamic Mechanical Properties
3.10. Thermophysical Properties and Morphological Features of Superconducting Polymer–Ceramic
Nanocomposites
3.11. Effect of Filling Fraction on the Valence State of Copper in Polymer–Ceramic Nanocomposites
3.12. Aging of Y1Ba2Cu3O6.97 Ceramics and Nanocomposites Containing Different Binders
3.13. Superconducting Polymer–Ceramic Nanocomposites Produced by GasPhase Ethylene Polymeriza
tion and Their Aging
3.14. Aging of Superconducting Polymer–Ceramic Nanocomposites Produced by the Frontal Polymeriza
tion of Metal Complex Monomers in the Presence of Ceramic Y1Ba2Cu3O6.97

1. INTRODUCTION
Discoveries related to nanomaterials have led to a true revolution in science and technology, and many
fundamental studies have become important for practical applications, having covered almost all areas of
research and found their niche in the hightechnology sector. It is valid to say that polymer nanotechnol
ogies occupy a special place in the field of nanotechnologies. The reason for this is that there is a growing
need for polymer nanocomposites in all of the areas of human activity, which is due to the wide possibili
ties of controlling the properties of polymer nanocomposites by varying not only the polymer matrix but
also the nature, dimensions, and properties of the nanofiller. It is therefore not surprising that there is
immense interest in the synthesis of such composite materials. Note that those studies are coming to the
forefront which employ results of basic research to develop technologically attractive processes for the
synthesis of polymer nanocomposites for a wide range of applications. Currently, a priority issue is the
development of commercially viable processes for the fabrication of polymer nanocomposites with tai
lored properties, which requires detailed knowledge of synthesis kinetics, underlying mechanisms, and
the properties of resulting products with allowance for the specifics and problems of the practical imple
mentation of processes for the synthesis of polymer nanocomposites.
The unique, sizedependent properties of nanoparticles pose serious problems when they are used as
fillers or stored and even in physical characterization processes. Having an extremely large surface area
and, accordingly, high surface energy, nanoparticles tend to stick together, forming large aggregates when
treated in solution or employed in practical applications. The most serious drawback of nanoparticles is
that they aggregate (agglomerate) when used as fillers in polymer composites: they stick together to form
large agglomerates not only during polymerization but also in the initial monomer medium, long before
the polymerization process. Naturally, this leads not only to the loss of the extremely valuable, special
properties of nanoparticles but also to an uneven distribution of the agglomerated, increased particles, so
that not only does the composite lose the properties expected from added nanoparticles, but it also has a
nonuniform distribution of the disperse phase over the polymer matrix. In this context, polymer nanotech
1 nology pays particular attention to issues pertaining to deagglomeration and a uniform filler distribution over
both the monomer medium and nanocomposite. Thus, it is the properties underlying the advantages of
nanoparticles and related materials which pose serious problems and difficulties in nanotechnology.
1 An important role in resolving the problem of nanofiller deagglomeration and uniform nanofiller dis
tribution in nanocomposites is played by frontal polymerization owing to the specifics of this process:
autooscillating polymerization mechanism. This review will discuss details of frontal polymerization
with application to nanotechnologies.
Many scientific disciplines (colloid chemistry, plasma polymerization, particle stabilization and clad
ding with various polymers and membranes, etc.) that have hitherto been of little technological interest
have become leading and fundamental areas of chemistry in the context of the search for approaches to
the stabilization of ultrafine particles. In this connection, the Ostwald theory declared unfounded—which
deals with disperse systems and assumes the primacy of the particle size over the other properties of mul
tiphase disperse systems—acquires special importance from the viewpoint of the fabrication of both poly
mer composite materials and any other materials. We mean that one should not neglect important pro
cesses, such as adsorption, interaction between disperse filler and polymer binder particles, and, finally,
chemical interaction between them, and, all other conditions being the same and with allowance for the
above factors, should investigate the effect of the size of the disperse phase in powder additives on the prop
erties of polymer composite materials.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 3

A diversity of inorganic nanoadditives are used in the fabrication of hybrid polymer nanocomposites:
metals, metal oxides, ceramics, nanoparticulate carbon black, nanotubes, modified clays (bentonite,
montmorillonite, laponite, and others), and layered disperse fillers having nanopores or nanolayers.
The synthesis of polymer nanocomposites has been the subject of extensive studies, which have pro
vided detailed information about the physicomechanical [1–17], optical [18–30], luminescent [31–37],
catalytic [38–5r1], electrical [52–62] and other properties [63–66] of polymer nanocomposites.
Clearly, the properties of polymer nanocomposites are governed by interfacial effects that occur at the
nanoparticle–macromolecule interface, the nature of the filler–matrix interactions (physical and chem
ical), and the morphological features of the binder [12–14, 67–70] in the interfacial layer.
It is worth pointing out that the structure and morphology of the interfacial layer in polymer nanocom
3 posites containing semicrystalline binders (homopolymers and copolymers) can be probed rather effec
tively by electron microscopy and Xray diffraction, but these techniques provide little or no reliable infor
mation about the structure of the interfacial layer in amorphous binders. In view of this, the first part of
this review discusses results of studies in which various approaches (radical, emulsion, and microemulsion
polymerizations of methyl methacrylate, styrene, and other monomers under sonication, and frontal
polymerization of methyl methacrylate and acrylamide) were used to synthesize a variety of polymer
nanocomposites. We consider the thermophysical, thermochemical, and dynamic mechanical properties
of the synthesized materials and discuss in detail the formation of the interfacial layer of a rigid amorphous
fraction (RAF) on the surface of nanoparticles. The second part addresses issues pertaining to the synthe
sis, properties, and morphology of superconducting intercalated polymer–ceramic nanocomposites con
3 taining amorphous or semicrystalline binders.
The interest in superconducting polymer–ceramic nanocomposites is motivated, among other things,
by the fact that, after the discovery of highTc superconductivity by Bednorz and Muller [71, 72], extensive
research efforts have been concentrated on the technological aspects of the fabrication and shaping of
superconducting articles of various geometries, containing metallic [73–82] or polymer [83–114] binders.
The small coherence length, electricalcurrent anisotropy, the strong effect of oxygen content on the
superconducting state, and the atmospheric degradation of the grain surface in SC ceramics create serious
difficulties in this area of research. Because of this, the fabrication and characterization of long lengths of
oxide ceramic superconductors requires knowledge of the phase equilibria involved, the structure of the
phases present, and the valence state of the constituent elements of the ceramic and a fundamental under
standing of the chemical and physical phenomena underlying the interaction of ceramics with inorganic
and organic binders. This situation results primarily from the nontrivial interaction of individual elements
and constituents (especially in the case of polymer binders) with the surface of ceramic grains and the pos
sible intercalation of the elements into the layered structure of the ceramic.
In a number of studies [73–82], silver was used as a metallic binder [73, 79, 81, 82] in the fabrication
of currentcarrying highTc superconductors. Silver molecules present on pore surfaces in ceramic grains
were found to migrate rather rapidly [79, 81]. Grainboundary segregation in SC ceramics was also
observed in the case of Mo, Zr, and Srcontaining binders [77].
As pointed out above, several reports described polymer–ceramic composites, with thermoplastic
[84, 87–113] and thermosetting [86] materials as binders. Goto and Kada [84] performed polymer binder
burnout as the final step of the forming process. The burnout of the organic component of composites typ
ically leads to an irreversible decrease in the oxygen content of the ceramic and its amorphization because
5 of the distortion of the orthorhombic SC phase. This accounts for the complete degradation of the SC
properties after burnout and the need for a special step to restore them.
Tonoyan and Davtyan [96] utilized polymer additives to protect highTc superconductors from atmo
spheric moisture. In many studies [91–113], appropriate polymer and ceramic (Y1Ba2Cu3O7 – x) powders
were thoroughly mixed in an agate mill [91, 101–104] to give a homogeneous mixture. The mixtures thus
prepared were loaded into preheated (130, 150, 160, or 200°C) dies and pressed at 100 MPa for 5, 10, 20,
or 30 min.
To produce polymer–ceramic nanocomposites by hot pressing [91–113], a variety of polymer binders
have been employed [114–119]. All of the polymers had the form of fine powders. The antioxidants used
were Irganox and HG2246 in an amount of 5 wt % relative to the polymer binder. Two compositions of
SC Y1Ba2Cu3O7 – x ceramics were used [91–113], Y1Ba2Cu3O6.97 and Y1Ba2Cu3O6.92, with a superconduct
ing onset temperature (Тс) of 93 and 91.5 K and transition width ((ΔТ = Тс – Тf) 6) of 6 and 6.5 K, respec
tively.
The physicomechanical, thermophysical, thermochemical, dynamic mechanical, superconducting,
and other properties of polymer nanocomposites were studied by techniques described in detail in Refs.
[91–113].

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


4 DAVTYAN et al.

2. PARTICLEFILLED POLYMER NANOCOMPOSITES


2.1. PMMA/SiO2, PS/SiO2, PMMA/Clay, and PS/Clay Nanocomposites Produced by Different Techniques
and Their Characteristics
The type of mobility of polymer macromolecules is often characterized in terms of the classical con
cept of glass transition, which can be followed using calorimetry or another technique. Available macro
molecule mobility and glass transition data for polymer nanocomposites are not very reliable. The reason
for this is that the mechanism of nanoparticle surface modification in the polymer–filler interfacial layer
has not yet been studied in sufficient detail. For example, the glass transition was variously reported to shift to
higher [114–120] or lower [119, 121–123] temperatures, remain unchanged [114, 117, 119, 121, 124, 125], or
completely disappear [119, 125–127]. There is ample experimental evidence that the mobility of the
binder is limited by the nanoparticle–macromolecule interfacial interaction, which leads to the formation
of a thin polymer layer. The presence of such interfacial layers in polymer nanocomposites has been dem
onstrated by various techniques and for various nanofillers [121, 124, 128–134]. In several studies [124,
127, 130], the interfacial layer was thought of as an immobilized fraction of the polymer. In other studies,
it was fixed at high temperatures by a socalled secondary glass transition [135, 136]. A second peak observed in
mechanical loss ( tan δ ) curves is, as a rule, thought to be an alternative feature. It is of interest to note that such
secondary transitions are interpreted not as the glass transition of interfacial layers [135–138] but as the forma
tion of microscopic gels in nanocomposites. Clearly, the observed secondary peak in mechanical loss may
be due not only to the glass transition. The main approach for identifying a glass transition in a material is
to measure and analyze its specific heat capacity, which is not considered in the literature in relation to the
entropy change characterizing relaxation processes [139–141].
3 It is worth pointing out that, as shown using semicrystalline polyethylene terephthalate (PETP) as an
example, it is important to obtain experimental data by dynamic characterization techniques and deter
mine the rigid amorphous fraction (RAF) [142, 143].
Even calorimetric techniques (as a rule, DSC), which are commonly used to study nanocomposites,
allow one to measure only the glass transition temperature. In a number of studies, the behavior of the
glass transition temperature was examined [119, 124, 137, 144–148] and specific heat capacity was deter
mined quantitatively. Klonos et al. [146, 147], assessed the effect of SiO2 and TiO2 nanoparticles on the
thermophysical properties of polymer nanocomposites and the molecular mobility of binder macromole
cules (polydimethylsiloxane [146, 147] and natural rubber). Their results demonstrate that the segmental
relaxation of the RAF of the polymer formed on the surface of the nanoparticles is two to three orders of
magnitude slower than that of the mobile part of the polymer.
Interestingly enough, Lipatov et al. [117, 128] have long understood the importance of measuring
absolute specific heat as a key thermodynamic characteristic of polymer composites. They determined the
specific heat of composites of SiO2 particles and PMMA or PS as a binder. Note also that the RAF can be
3 determined by a method known in the literature for semicrystalline polymers, which was proposed by
Wunderlich [148, 149].
There are still unanswered important questions regarding the RAF devitrification (relaxation) temper
3 ature in semicrystalline polymers [150–152] and the sequence of RAF relaxation and polymer melting
processes.
The picture is much simpler in the case of polymer nanocomposites containing inorganic nanoaddi
tives. Indeed, in the temperature range where the polymer–nanoparticle interaction does not vary, inor
ganic nanoparticles (metals, their oxides, etc.) undergo neither melting no phase transitions. Because of
this, polymer nanocomposites containing SiO2 are better suited for studying the glass transition of the
RAF formed in the nanoparticle–polymer interfacial layer.
It is worth noting that the secondary glass transition, described in several reports [136–138], is also related
to the RAF of the nanocomposite. In all cases, however, the secondary glass transition is interpreted, based
on dynamic measurement results, as an individual peak, or its part, for the α relaxation process. At the same
time, it should be noted that no evidence of a secondary glass transition in polymer nanocomposites was
found in special calorimetric studies. This led several groups [12, 13, 67–69, 146, 147] to synthesize polymer
nanocomposites containing an appreciable rigid amorphous fraction in order to accurately determine
their specific heat, to identify a secondary transition from an increase in specific heat as a criterion, and
evaluate the segmental mobility of the polymer in the form of an RAF [146, 147].
For such studies, PMMA/SiO2, PS/SiO2, laponite RDTM/SiO2 (TiO2), and natural rubber/SiO2 (TiO2)
nanocomposites were produced using emulsion and microemulsion polymerizations of ST and MMA,
frontal polymerization of MMA, and sol–gel processing, and their thermophysical, dynamic mechanical,

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 5

2 μm

Fig. 1. Electron micrograph of a dried PMMA/laponite RD nanocomposite (microemulsion polymerization, 11 wt %


laponite RD).

and thermochemical properties, the formation of the RAF, and its segmental mobility were investigated
[12, 13, 67–69, 146, 147]. Consider those studies in greater detail.

2.2. Microemulsion Polymerization


In the preparation of PMMA/SiO2, PS/SiO2, and PMMA/laponite RD nanocomposites in recent
studies [12–14, 67–70], no additional compounds were introduced into the polymerization medium
wherever possible in order to avoid additional interactions between the polymer macromolecules and
nanoparticles. The experimental procedure was described in detail in Sargsyan et al. [67]. Here we note
only that, to avoid SiO2 nanoparticle agglomeration, the reaction medium was sonicated at 250 W for
20 min. The appearance of the polymer composite thus produced is illustrated in Fig. 1. The number
average molecular weight of the pure polymer was Mn = 6.1 × 104 and its polydispersity index was 5.4.
Of special note is that laponite RD contains an unstable, rapidly vaporizing phase (≈10%), which was
taken into account in weight loss measurements (Fig. 2) with a PerkinElmer thermoanalytical system [67].
After centrifugation and vacuum drying, the nanocomposites were pressed at 200 kPa and 150°C and
examined by transmission electron microscopy (TEM). Figures 3a and 3b show micrographs of the
PMMA/laponite RD nanocomposites at different magnifications. As seen in Figs. 3a and 3b, the laponite
particles reside mostly on the latex surface. They are wellresolved. At a filling fraction of 11 wt %, TEM
demonstrates that the material has a layered structure (Fig. 3b).
A natural question in this context is what amount of the filler will ensure that the clay surface will be
covered with polymer binder macromolecules? Given that the laponite RD particles have a diameter of
150 nm, thickness of 1 nm, and density of 2.4 g/cm2 and that the density of the polymer is 0.936 g/cm2,
we find that the surface layer of the laponite RD will be coated with the polymer at a filler content of
10 wt %. It is of interest to note that this estimate agrees well with what is seen in Fig. 3.
In the case of spherical SiO2 nanoparticles in PMMA/SiO2 nanocomposites, a small amount of
agglomerates forms (Fig. 4).

% weight loss % weight loss % weight loss


0 9 0 7 0 5
20 8 20 20
7
40 40 6 40
6 4
60 60 5 60
(a) 5 (b) 4 (c)
80 4 80 3
3 3 80
2 2 2
100 1 100 1 100 1
100 200 300 400 500 600 100 200 300 400 500 600 100 200 300 400 500 600
Temperature, °C Temperature, °C Temperature, °C
Fig. 2. Weight loss of (a) PMMA/SiO2, (b) PMMA/laponite RD, and (c) PS/SiO2 nanocomposites.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


6 DAVTYAN et al.

(a) 500 nm (b) 50 nm

Fig. 3. (a, b) TEM micrographs of PMMA/laponite RD nanocomposites. The samples were prepared by microemulsion
polymerization at 150°C and 200 kPa [67].

10 nm

(a) 200 nm (b) 50 nm

Fig. 4. TEM micrographs of PMMA/SiO2 nanocomposites (4 wt % SiO2) at different magnifications.

2.3. Formation of the Immobilized Polymer Fraction


3 It is known that, rather frequently, the specific heat Δ C p of semicrystalline polymers changes in
response to their glass transition much less than might be expected. It is for this reason that Wunderlich
introduced the concept of rigid amorphous fraction (RAF) [148, 149]:
RAF = 1 – degree of crystallinity – Δ C p Δ C p pure , (1)
3 where Δ C p and ΔC p pure are the specific heat changes for a semicrystalline and an amorphous polymer,
respectively, at their glass transition temperatures. The ratio of these quantities represents the fraction of
the polymer that is involved in the glass transition and is referred to as the mobile amorphous fraction
(MAF). As shown earlier [143, 144, 153], Eq. (1) can be used to assess the relative specific heat in dynamic
glass transitions. In the process of dynamic relaxation, the RAF was determined using dielectric, dynamic
mechanical, and DSC measurements. It is of interest to note that according to dielectric and mechanical
3 loss data, Eq. (1) is unsuitable for determining the RAF of semicrystalline polyethylene terephthalate in
secondary, more local, βrelaxation processes. The point is that βrelaxation processes are observed in
3 both the mobile and immobilized amorphous fractions of semicrystalline polymers, but not in their crys
talline part, as was demonstrated in several studies [143, 144, 153]. In those studies, the assumption was
made that the RAF is immobilized because no cooperative motion of binder macromolecules was
detected. The characteristic cooperative mobility was assessed by calorimetry [140, 154, 155] at a scan step
of several nanometers.
The decrease in relative specific heat Δ C p as a result of the glass transition in polymer nanocomposites
was discussed by Yagrarov et al. [134]. Clearly, the MAF or RAF can be determined using (1) after the
degree of crystallinity is replaced by the number of nanoparticles [114, 138]:
RAF = 1 – filler content – Δ C p Δ C p pure (2)
To apply Eq. (2) to polymer nanocomposites in a vitreous state, precise temperaturedependent spe
cific heat data are necessary. The point is that, even after careful drying of nanocomposites at temperatures
above their glass transition temperature (Tg) or in vacuum, the samples may contain residual solvents or

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 7

9
2.2 (a) 8
2.2 (b) 7
7 6
2.0 6 2.0 5
1.8 5 4
4 1.8 3
1.6 1.6 2
1.4 3
2 1.4
1.2 1.2

Specific heat, J/(K gsample


1.0 1 1.0 1
SiO2 SiO2
0.8 0.8
100 60 80 100 120 140 160 100 60 80 100 120 140 160
Temperature, °C Temperature, °C
2.2 (c) 5
2.0 4
1.8 3
1.6 2
1.4
1.2
1.0 SiO 1
2
0.8
100 60 80100 120 140
Temperature, °C

Fig. 5. Temperature dependences of specific heat for (a) PMMA/SiO2 (b) PMMA/laponite RD, and (c) PS/SiO2 nano
composites. The unnumbered curves refer to pure PMMA and PS and were obtained using the ATHAS scheme [158] and
step scans. The binder : nanoadditive weight ratio is (a) (1) 100 : 0, (2) 96 : 4, (3) 85 : 15, (4) 78 : 22, (5) 70 : 30, (6) 53 : 47,
(7) 34 : 66, (8) 27 : 73, (9) 0 : 100, (b) (1) 100 : 0, (2) 89 : 11, (3) 86 : 14, (4) 73 : 27, (5) 58 : 42, (6) 41 : 59, (7) 0 : 100,
(c) (1) 100 : 0, (2) 91 : 9, (3) 76 : 24, and (4) 54 : 46.

Tg sample−Tg pure, K
8
7 3
6
5 2
4
3 1
2
1
0
−1
0 10 20 30 40 50 60 70
Percent assitive

Fig. 6. Glass transition temperature as a function of filling fraction for (1) PMMA/SiO2, (2) PMMA/laponite RD, and
(3) PS/SiO2 nanocomposites. The Tg of PMMA and PS are 117 and 99°C, respectively.

other substances, which will vaporize during specific heat measurements. To obtain more reliable data,
Sargsyan et al. [67–69] performed precision specific heat measurements, taking into account possible side
processes (vaporization [156]). Prior to DSC scans, nanocomposite samples were heated to 170°C at a rate
of 20 K/min and then quenched in 10 min. To improve sample purity control, dry pure nitrogen was used
[67–69] and the samples were weighed after the measurements.
Figure 5 shows the temperature dependences of specific heat for PMMA and PSmatrix nanocom
posites.
The curves in Fig. 5 were obtained with allowance for the sample (polymer + nanoadditive) weight.
Because of the low specific heat of SiO2 nanoparticles, increasing their content in the nanocomposites
3 reduces its specific heat. Clearly, such changes cannot be observed in semicrystalline polymers because, at
temperatures below the glass transition temperature, the RAF and MAF differ little in specific heat. Inde
pendent of the nature of the binder and nanofiller, the glass transition temperature increases slightly in the
case of the PMMA/SiO2, PMMA/laponite RD, and PS/SiO2 nanocomposites (Fig. 6) [67].

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


8 DAVTYAN et al.

ΔCp sample−ΔCp pure


1.0
0.9 Nanoadditive
0.8
0.7
0.6 RAF
0.5
0.4 2 1
0.3
0.2 3
MAF 4
0.1
0
0 10 20 30 40 50 60 70 80 90 100
Percent assitive

Fig. 7. Relative specific heat as a function of nanoadditive content for (2) PS/SiO2, (3) PMMA/SiO2, and (4)
PMMA/laponite RD [67]. The lower vertical twoheaded arrow represents the RAF at a laponite RD content of 35 wt %.

5
2.3 (a) 7 2.3 (b) 4
5−6 3
4 2
2.1 3 2.1 1
1.9 2 1.9
1
1.7 1.7
1.5 1.5
Specific heat, J/(K gsample

1.3 1.3
40 60 80 100120140160 40 60 80 100120140160
Temperature, °C Temperature, °C
2.1 (s) 32
1
1.9
1.7
1.5
1.3
1.1
40 60 80 100 120 140
Temperature, °C

Fig. 8. Temperature dependences of specific heat for (a) PMMA/SiO2 (b) PMMA/laponite RD, and (c) PS/SiO2 nanocom
posites. The lines with no data points were obtained for PMMA and PS using the ATHAS scheme [158]. The binder :
nanoadditive weight ratios are the same as in Fig. 5.

3 This effect differs from that in semicrystalline polymers, where the crystallization process typically
leads to a considerable increase in their glass transition temperature and extends their glass transition
range [158]. It follows from the data in Fig. 5 that the specific heat decreases with increasing nanofiller
content (Fig. 5a, curves 1–8; Fig. 5b, curves 1–6; Fig. 5c, curves 1–4).
Figure 7 plots the relative specific heat against nanoadditive content for PS/SiO2 (line 2), PMMA/SiO2
(line 3), and PMMA/laponite RD (line 4) nanocomposites. Line 1 in Fig. 7 represents the additivity rule
specific heat values for the nanocomposites.
It follows from line 2 in Fig. 7 that the data obtained for the PSmatrix nanocomposites are similar to
line 1, whereas the relative specific heat of the PMMAmatrix nanocomposites varies in a more complex
manner (lines 3, 4).
The relative specific heat decreases more rapidly at small amounts of nanoadditives and then varies in
parallel with line 1. The parallel portion begins at ~30 wt % additive. Such behavior of the relative specific
heat (line 3) can be interpreted as evidence that there are rather strong bonds between the surface of the

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 9

1.0
0.9 1

Filler
0.8

RAFpol
ΔCp sample/ΔCp pure
0.7
0.6 2
0.5
0.4
0.3 3

MAFpol
0.2
0.1
0
0 10 20 30 40 50 60 70 80 90 100
wt % filler

Fig. 9. Relative specific heat as a function of nanofiller content for the polymer fraction in (1) PS/SiO2, (2) PMMA/SiO2,
and (3) PMMA/laponite RD. The vertical arrows represent the RAFpolymer and MAFpolymer in the PMMAmatrix nano
composites containing 66% SiO2.

1.0
14 1

1'
0.8
13

Specific heat,
J/(K gsample)
(upward = endo)
Heat flux, mV

2 0.6
12
2'
11 0.4
10
0.2
3 3'
9
0
8
70 80 90 100 110 120 130 140 150
Temperature, °C

Fig. 10. Heat flux and specific heat as functions of temperature for PMMA. Heat treatment at 105°C for 10 h. (1, 2) heat
ing after heat treatment, (1', 2') heating without heat treatment, (3, 3') specific heat of pure PMMA (24 mg) and a
PMMA/SiO2 nanocomposite (25 mg) containing 35 wt % SiO2 [67–70].

SiO2 nanoparticles and the polymer macromolecules, which lead to the formation of a RAF in the poly
mer–nanoparticle interfacial layer. The RAF differs drastically in properties from the bulk polymer. Most
likely, increasing the nanoadditive content (to 25–30% and above) leads to a marked increase in particle
size through agglomeration, thereby reducing the specific surface area of the resulting agglomerates and,
accordingly, the surface tension energy.
Because of this, at filler contents above 25–30% the relative specific heat decreases in parallel with line 1.
The behavior of the relative specific heat of the PMMA/laponite RD nanocomposites can be explained in
a similar way.
The initial portions of lines 1, 3, and 4 correspond to a constant RAF : nanofiller ratio. Most likely, this
situation corresponds to an ideal state: the nanoparticles are coated with identical amounts of RAF, which
depend on the geometric shape, size, and nature of the nanofiller and monomer.
Clearly, the large difference between lines 3 and 1, as well as that between lines 4 and 1, indicates that
there is a marked difference in mobility between the PMMA molecules in the RAF and MAF states. At the
same time, Fig. 7 demonstrates that, in the case of the PS/SiO2 nanocomposites, lines 1 and 2 differ very lit
tle. Indeed, according to the data in Figs. 8a and 8b, the PMMAmatrix nanocomposites differ markedly,
independent of the nature of the binder, especially around their glass transition temperatures. Figs. 8a and
8b clearly show that, at the glass transition temperatures, a decrease in nanofiller content leads to an
increase in specific heat jump, which can be accounted for by RAF formation in the PMMAmatrix nano
composites. At the same time, varying the amount of SiO2 in the PSmatrix composites has no effect on
the specific heat jump near their glass transition temperatures (Fig. 8c).
To describe lines 3 and 4 in Fig. 7 over the entire range of nanofiller contents studied, it is convenient
to use Eq. (2) in a modified form [11–13]:

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


10 DAVTYAN et al.

Enthalpy change, J/gpol

1.2

0.8 1
2
0.4

−120 −80 −40 0 40


Tfiring – Tg , K

2 Fig. 11. Enthalpy change (J/gpolymer) after heat treatment for 10 h as a function of the difference between the heat treat
ment and glass transition temperatures for pure PS and PS/SiO2 nanocomposites containing 24 wt % SiO2 [67–70].

ΔH, H/gpol

1.2

0.8

0.4 1

2
0

−120 −80 −40 0 20


Tfiring – Tg , K

2 Fig. 12. Enthalpy change (J/gpolymer) after heat treatment for 10 h as a function of the difference between the heat treatment
and glass transition temperatures for pure PMMA and PMMA/SiO2 nanocomposites containing 35 wt % SiO2 [67–70].

dnm
3.0

2.6

2.2

1.8 2
1.4

1.0 1
0.6
0 10 20 30 40 50 60 70 80
Percent additive

Fig. 13. Thickness of the RAF as a function of nanofiller content in (1) PMMA/SiO2 and (2) PMMA/laponite RD nano
composites [67–70].

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 11

Specific heat, J/(K gsample)

2.6
2.4
2.2
2.0
1.8
1.6
1.4
60 100 140 180 220
Temperature, °C

Fig. 14. Specific heat versus temperature data for pure PMMA and the PMMA/SiO2 nanocomposite containing 47 wt %
filler. Step scan DSC; sample weight, 21 mg. The straight lines were obtained for PMMA using the ATHAS scheme [157].

Specific heat, J/(K gsample)


2.6
2.4
2.2
2.0
1.8
1.6
1.4
50 100 150 200 250 300 350
Temperature, °C

Fig. 15. Specific heat data obtained during heating at 400 K/m for pure PMMA and the PMMA/SiO2 nanocomposite
containing 47 wt % filler. Step scan DSC; sample weight, 0.5 mg. The straight lines were obtained for PMMA using the
ATHAS scheme [157].

⎧1 − ε − ΔC p sample / ΔC p pure, for ε < ε cr


MAF = ⎨ (3)
⎩1 − ε − [ΔC p sample / ΔC p pure]cr , for ε ≥ ε cr
where ε is the filler content before agglomeration, and εcr and [ΔCp sample/ΔCp pure]cr are the filler content
and relative specific heat at the beginning of active agglomeration. Clearly, both εcr and
[ΔCp sample/ΔCp pure]cr depend on the nature, geometric shape, and size of the nanoparticles.
In Fig. 8, the specific heat step was evaluated at the glass transition temperature, as is common prac
tice, and was then determined more accurately using specific heat data for the pure polymer. The accuracy
in the step heights obtained depends on the tangent to the curves in Fig. 8 at the inflection point. A key
point is to accurately determine the specific heat ΔCp polymer above the glass transition temperature.
When accurate data are obtained above the glass transition temperature, the curves in Figs. 8a and 8b
should be parallel, independent of nanofiller content. Based on the conclusions drawn previously [67], we
determined the RAFpolymer from the data in Figs. 7 and 8 for PMMA/SiO2, PMMA/laponite RD, and
PS/SiO2 nanocomposites. The results are presented in Fig. 9.
Comparison of Eq. (3) with the data in Figs. 7 and 9 provides no evidence for an immobilized fraction
(RAF) in the PS/SiO2 nanocomposites (line 1).
Thus, the data in Figs. 7 and 9 indicate with certainty that the PMMA/SiO2 and PMMA/laponite RD
nanocomposites contain a RAF, whereas no RAF forms in the PSmatrix nanocomposites. To ascertain
that the PS/SiO2 nanocomposites contain no RAF, additional determinations are needed.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


12 DAVTYAN et al.

(a) (b) (c)

Fig. 16. Texture of a parent reaction medium in the AAm + water + FeO system: (a) no surfactant, (b, c) with a surfactant;
(c) 30×.

(a) (b)

Fig. 17. (a) Texture of a parent reaction medium in the AAm + water + YBa2Cu3O7 – x system with a surfactant; (b) the
same system at a magnification of 30×.

(a) (b) (c)

Fig. 18. Effect of the thermal conditions of the process on the structure of nanocomposites prepared by AAm polymer
ization: (a) adiabatic, (b) steadystate frontal, and (c) oscillating frontal polymerization.

As discussed above, to ascertain the presence of a RAF, one should determine Δ C p along tangents to
heat capacity curves (Figs. 5, 8). Note in this context that, as seen from the shape of the curves in Figs. 5
and 8, in some cases it is impossible to draw exact tangents to specific heat curves, which is the obvious
reason for the fact that the RAF cannot be accurately determined.
At glass transition temperatures and below, the molecular mobility of polymer nanocomposites can be
checked by heattreating them [159]. If the amorphous fraction of the polymer in a polymer nanocom
2 posite is immobile, its enthalpy should decrease below its glass transition temperature [160–162]. To val
idate this assumption, Sargsyan et al. [67–69] carried out special heat treatment of their samples. The
samples were heated to above the glass transition temperature (170°C) and then cooled to 105°С at a rate

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 13

Δcp sample/Δcp pol

1.0

0.8
3
0.6
2
0.4 RAF
1
0.2

0 20 40 60 80 100
wt % SiO2

Fig. 19. Plot of ΔCp sample/ΔCp pol against the percentage of SiO2 added to the polymerization medium. Particle size of
(1) 10 nm, (2) 0.6 μm, and (3) 30–50 μm.

of 10 K/min and held there for 10 h. After this heat treatment, the samples were cooled to 30°С and again
2 heated to 170°С in order to determine the enthalpy change.
A characteristic heat treatment peak was observed during the first heating (Fig. 10, curve 1) and was
missing during heating without heat treatment (Fig. 10, curve 1').
The difference between these two curves can be used to assess the specific heat change as a function of
temperature (Fig. 10, curves 3, 4). Integrating curves 3 and 4 over the temperature range 80–149°C, we
2 find the enthalpy change.
2 To compare the enthalpy change in a pure polymer to that in a nanocomposite, we took into account
2 the polymer fraction in the nanocomposites. Figures 11 and 12 show the enthalpy change as a function of
the difference between the heat treatment (105°С) and glass transition temperatures for pure PMMA, PS,
PS/SiO2, and PMMA/SiO2.
2 As seen in Fig. 11, the enthalpy curve of pure PS is identical to that of the nanocomposite, in contrast
to what is seen in Fig. 12 for PMMA.
Thus, earlier results [67–70] lead us to conclude that PMMA/SiO2 and PMMA/laponite RD nano
composites contain a RAF, in contrast to PS/SiO2 nanocomposites (Figs. 7, 10, 11). The question that
arises in the context of these results is what is the thickness of the RAF formed on the surface of the nano
filler particles? For spherical SiO2 nanoparticles of 10 nm average diameter, a laponite RD thickness of
1 nm, polymer density of 1 g/cm3, and SiO2 nanoparticle density of 2.4 g/cm3 [67–69], we obtain an RAF
thickness of ~2–2.5 nm, which follows directly from (3) and the data in Figs. 7 and 9. It is worth pointing
out that the RAF in the PMMA/laponite RD nanocomposites is considerably greater than that in
PMMA/SiO2. Interestingly enough, before active agglomeration, when the filler content is within 30 wt %,
the thickness of the RAF (Fig. 13) for the SiO2 and laponite RD fillers is ~2 and ~2.5 nm. Note that the
3 thickness of the RAF is ~2 nm in semicrystalline polyethylene [155] and ~1.5 nm in rubber [164].
It is seen in Fig. 13 that, at filler contents under 25–30% in the PMMA/SiO2 nanocomposites and
under 35–40% in the PMMA/laponite RD nanocomposites, the thickness of the RAF layer is constant.
At higher filler contents, it decreases. The behavior of curves 1 and 2 in Fig. 13 can be accounted for by
the active agglomeration processes at high filler contents.
It is known that, in the interfacial layer within corearranged regions (CRRs) [140, 154, 163, 164],
some of the immobilized polymer macromolecules influence the segmental mobility of neighboring
chains and are responsible for liquidlike motions. It is worth pointing out however that the entire CRR
cannot contribute to liquidlike motions and, hence, cannot increase the specific heat at the glass transi
tion temperature.
If the formation of an immobilized polymer fraction on the surface of nanoparticles is the result of
polymer–nanoparticle interaction, one should obviously expect significant changes in the devitrification
3 process. The statement that the RAF of semicrystalline polymers devitrifies only gradually and above the
main glass transition temperature [149] is disputable. Under the assumption that local polymer chain

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


14 DAVTYAN et al.

40 nm 500 nm 1 μm
(a) (b) (c)

Fig. 20. (a, b) Transmission and (c) scanning electron micrographs illustrating (a) a uniform distribution and (b, c)
agglomeration of SiO2 nanoparticles in a polymer matrix.

immobilization on the surface of nanoparticles is due to a rather thick RAF layer, it is reasonable to expect
that the devitrification of this layer is necessary to completely eliminate the corresponding interaction.
The specific heat will not gradually increase as long as there are the mentioned interaction forces in the
interfacial layer. It is therefore reasonable to expect that zero polymer–nanoparticle interaction will allow
the immobilized fraction to devitrify.
To examine the behavior of the RAF layer and a possible secondary glass transition, Sargsyan et al.
[67–69] measured the specific heat change down to the thermooxidative destruction temperatures of the
polymer using DSC step scans. The temperature was varied in 5K steps, and precise heat capacity mea
surements were made under isothermal conditions. Figure 14 presents specific heat versus temperature
data for the PMMA/SiO2 nanocomposites containing 47 wt % filler.
It might be expected that RAF devitrification will show up as a secondary transition in specific heat.
It is, however, seen in Fig. 14 that no such transition was detected when the material was heated until the
polymer destruction onset, that is, to 250°C.
To reduce the effect of polymer destruction on the specific heat, samples were heated at 400 K/min
[67–69], which was expected to markedly increase the thermal destruction onset temperature. Indeed, it
is seen in Fig. 15 that, during heating at a rate of 400 K/min, the polymer in the nanocomposites does not
decompose up to ~360°C, but no secondary glass transition is seen either.
This result suggests that there is a rather strong polymer–nanoparticle interaction, which persists at
high temperatures.
It is worth pointing out that, before these results were obtained [67–70], the interaction between
PMMA macromolecules and the surface of SiO2 nanoparticles had been assumed to be weaker than in the
3 case of covalent bonds, like in semicrystalline polymers [151].

tanδ
E'' ⋅ 10−3, MPa (a) 4 (b) 4
4 3
15 3 3
2
10 2
1 2
5 1
1

0 0
20 60 100 140 180 20 60 100 140 180
Temperature, °C Temperature, °C

Fig. 21. Temperature dependences of the (a) dynamic elastic modulus and (b) mechanical loss tangent. SiO2 content:
(1) 0, (2) 20, (3) 30, and (4) 45 wt % relative to the monomer.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 15

ν, cm/min

0
0.2 0.4 0.6 0.8 1.0 1.2 1.4
I0, wt %

Fig. 22. Effect of initiator (AIBN) concentration on the velocity of frontal polymerization of an AAm–bentonite mixture
at a filler content of 30%.

2.4. PAAM Nanocomposites Prepared under Different Thermal Conditions


To assess the nanoparticle distribution over the reaction system and resultant polymer matrix, Davtyan
et al. [13] and Tonoyan et al. [165] carried out a series of experiments with and without surfactant addi
tions. In the former case, FeO nanoparticles (10 wt % relative to the monomer) were added to an AAm–
water mixture. After stirring, samples were examined on a LYuMAM3 microscope.
Figure 16a is a micrograph of one of the AAm–water (20%) + FeO (10%) samples. It clearly demon
strates an uneven distribution of the FeO nanoparticles over the reaction system.
1 The addition of a surfactant (5%) to the above system leads to deagglomeration and a nonuniform dis
tribution of the nanoparticles over the system (Fig. 16b). Examination of the mixture at higher magnifi
cation indicates the formation of typical micellar structures (Fig. 16c), due to the interaction of the sur
factant with the nanoparticle surface. Note that increasing the temperature of the reaction medium to 60–
70°С distorts the uniform nanoparticle distribution and leads to agglomeration, and the system returns to
essentially the same state as it was before the addition of the surfactant (the micrographs are similar to that
in Fig. 16a and are not presented). The effect can be accounted for by the breakdown of the micellar struc
tures with an increase in temperature and, accordingly, agglomeration of the nanoparticles. The same
compositions as in the FeOcontaining system were used in studies of the AAm + YBa2Cu3O7 – x + water
system with surfactant additions. Figure 17a is an optical micrograph of a reaction medium containing
fine YBa2Cu3O7 – x powder (the same magnification as in Fig. 16a).
Examination at higher magnification (Fig. 17b) shows that typical micellar structures, of various sizes,
form in this reaction system as well. The scatter in micelle size in Fig. 17b is most likely determined by the
perovskite ceramic particles, which form cores, and the fine ceramic particles are aggregates of micellar
structures.
AAm containing evenly distributed micellar structures was polymerized under various thermal condi
tions: isothermal, adiabatic, and frontal (DCPD initiator concentration, 3 × 10–3 mol/L). The resultant
nanocomposites were examined under a luminescence microscope. The optical micrographs in Fig. 18
demonstrate that the polymer nanocomposites have various structures, depending on thermal conditions.
Adiabatic polymerization (Fig. 18a) is accompanied by nanoparticle agglomeration. As a result, the
filler distribution over the polymer matrix becomes nonuniform, which is due, as noted above, to the
breakdown of micellar structures.
Interesting structures were obtained by frontal polymerization. In the case of steadystate front motion
(no heat transfer from the reaction zone to the environment), a uniform nanoparticle distribution over the
polymer matrix was observed (Fig. 18b). The reason for this is that the heat wave of the polymerization
front fixes the nanoparticle distribution, eventually leading to the formation of a polymer nanocomposite
with a uniform particle distribution over the polymer matrix. Heat losses result in the formation of typical
layered structures (Fig. 18c), which is most likely due to reaction front velocity oscillations near a steady
state value [113, 166–168].
Isothermal AAm polymerization leads to an uneven distribution of FeO nanoparticles over the polymer
matrix. It is reasonable to assume that heating the reaction medium to the melting point of the crystalline
monomer (~75°С) results in the breakdown of micellar structures and, accordingly, agglomeration of the
FeO nanoparticles. Thus, there is special interest in frontal polymerization, which allows the initial dis
tribution of nanoparticles in the monomer to be fixed in a polymer matrix. In connection with this, the

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


16 DAVTYAN et al.

n n
(a) (b)
0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 10 20 30 40 0 10 20 30 40
Percent additive Percent additive

Fig. 23. Reaction order with respect to the initiator as a function of filler content for the frontal polymerization of AAm–
bentonite mixtures with (a) BP and (b) AIBN as initiators; ( ) descending and ( ) ascending waves.

polymerization behavior of MMA in the presence of SiO2 nanoadditives was studied in detail in frontal
mode [13]. As shown below, frontal MMA polymerization ensures a uniform nanoparticle distribution
even with no surfactant.

2.5. Formation of the Rigid Amorphous Fraction under Frontal Polymerization Conditions
and Thermophysical Characteristics of the Resultant Nanocomposites
In the preceding sections, we analyzed earlier results [67–70, 118, 119, 146, 147] that demonstrate that
the relative specific heat of nanocomposites produced by various techniques (emulsion and microemul
sion polymerizations and sol–gel processes) varies nonadditively. This was accounted for by the rather
strong interaction between the polymer chains and nanoparticle surface, which leads to RAF formation.
The RAF increases the glass transition temperature of the nanocomposites by 5–6°С.
It is of interest to compare the formation and dimensions of the RAF for filler particles of the same
nature but different sizes. To this end, consider results reported by Davtyan et al. [13]. Using frontal poly
merization, they synthesized nanocomposites containing SiO2 additions (with average particle sizes of
10 nm, 0.6 μm, and 30–50 μm) and investigated their thermophysical properties. Figure 19 plots their rel
ative specific heat against added filler content [13].
It follows from line 1 in Fig. 19 that the addition of SiO2 with an average particle size of 30–50 μm leads
to additive variation of the relative specific heats of the polymer and SiO2. This can be interpreted as evi
dence that there is no chemical interaction between the filler surface and PMMA macromolecules or that
the contribution of such interaction is insignificant, which is more likely.
Reducing the filler size to 0.6 μm leads to a change in the shape of the plot of ΔCp sample/ΔCp pol against
filler content (Fig. 19, line 2). The ΔCp sample/ΔCp pol ratio decreases more rapidly in comparison with line 3,
and only at filler contents of 15–20 wt % and higher is line 2 parallel to line 3. When nanoparticles (10 nm)
are used, the dependence of ΔCp sample/ΔCp pol on the percentage of added nanoparticles is identical to the
data in Fig. 7. The variation of ΔCp sample/ΔCp pol with filler content at particle sizes of 10 nm and 0.6 μm
(lines 1 and 2) suggests that there is rather strong interaction between the polymer binder macromolecules
and nanoparticle surface. The observed smaller deviation of the plot of relative specific heat against
nanoadditive content (Fig. 19) at an average particle size of 0.6 μm is probably due to the low content of
nanoparticulate SiO2 in the starting powder.
Clearly, lines 1 and 2 in Fig. 19 can be quantitatively represented using relation (3).
Relation (3) and the data in Fig. 19 can be used to determine the characteristic size (layer thickness) of
the RAF. Proceeding like in Section 2.2, we find that the thickness of the RAF is ~2–2.2 and ~0.5–0.7 nm
at an average SiO2 particle size of 10 nm and 0.6 μm, respectively.
That the RAF has identical thicknesses in the nanocomposites prepared by frontal and microemulsion
(under sonication) MMA polymerization in the presence of SiO2 (particle size, 10 nm) indicates that, like
1 highfrequency acoustic fields, a heat wave leads to particle deagglomeration, and the polymerization
zone, which follows the heating zone, fixes this state. Clearly, at high contents of SiO2 nanoparticles
1 (above 25–30%), a heat wave causes only partial deagglomeration of enlarged particles.
As pointed out above, the parallel portions of lines 1–3 in Fig. 19 are due to agglomeration of the nano
particles, which thus become nonreactive with the polymer matrix. To validate this conclusion, nanocom

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 17

posites differing in SiO2 content (10nm particles) were examined on a transmission electron microscope
[13] (Fig. 20).
It is seen in Fig. 20a that, at filler contents below 25–30%, the SiO2 nanoparticles are uniformly dis
tributed over the polymer matrix.
Increasing the nanofiller content to above 25–30% leads to particle agglomeration (Fig. 20b). Above
30–35% SiO2, the agglomerate size increases further (Fig. 20c).

2.6. Dynamic Mechanical Properties of PMMA/SiO2 Nanocomposites


Figure 21 shows the temperature variations of the dynamic elastic modulus (E') and mechanical loss
tangent ( tan δ ) for PPMA/SiO2 nanocomposites differing in SiO2 content (10nm particles).
Analysis of the temperaturedependent Е' data in Fig. 21a shows that increasing the nanofiller content
to 25–30% (curves 1–3) increases the dynamic modulus, whereas above 30% Е' is essentially independent
of SiO2 content (curve 4). The accompanying increase in glass transition temperature (curves 1–3) points
to RAF formation on the surface of the SiO2 particles. These results are consistent with data on the effect
of SiO2 content on the mechanical loss tangent. Indeed, it follows from the data in Fig. 21b (curves 1–3)
that the temperature range of the largest mechanical loss is associated with the devitrification of the poly
mer binder in the nanocomposite. An increase in SiO2 content (to 25–30%) is accompanied by broaden
ing of the mechanical loss peaks, a shift of the peaks to higher temperatures, and an increase in peak
height. As noted above, the observed increase in peak height and the shift of the peaks to higher tempera
tures are due to the rather strong interaction of the binder macromolecules with the surface of the filler
grains, which leads to RAF formation. At the same time, the origin of the secondary transition between
~140 and 190°С in the mechanical loss curves for the nanocomposites with a filler content of 30% and
above (Fig. 21b) is unclear. As mentioned above, a similar secondary transition as a component of the main
mechanical loss peak was observed by Tsagaropoulos and Eisenberg [136, 137] and Fragiadakis et al. [138]
and was attributed to a secondary glass transition of the polymer nanocomposites. The secondary transi
tion in Fig. 21b can probably be accounted for by the fact that individual macromolecules involved in RAF
formation may bind SiO2 nanoparticles, leading to the formation of structures resembling threedimen
sional networks [169]. In such structures, SiO2 particles may serve as network nodes. Clearly, such struc
tures can have various topologies.
The assumption that RAF formation is accompanied by the formation of macroscopic gels allows us to
account for the main and secondary mechanical loss transitions in Fig. 21b. In the range 100–150°С,
where the main transition occurs, we observe the devitrification onset of the polymer, including the poly
mer binder present in the structure of the RAF. At higher temperatures, in the range of the secondary tran
sition, the molecular mobility of the loose structures of the macroscopic gels manifests itself. Clearly, the
formation of threedimensional structures with SiO2 nanoparticles as nodes [13] may have a rather strong
effect on the properties of the synthesized polymer nanocomposites. Thus, the present results lead us to
conclude that frontal polymerization is a technologically advantageous factor that contributes to nanopar
1 ticle deagglomeration in a monomer medium and helps to maintain a uniform particle distribution in the
starting reaction medium and resulting polymer composite. As a result, there is a rather strong interaction
of the PMMA macromolecules with the surface of the nanoparticles. Accordingly, an RAF forms in the
interfacial layer, leading to marked changes in the relative specific heat, glass transition temperature, and
dynamic mechanical properties of the nanocomposites, depending on filler content.

2.7. Effect of Nanoadditives on the Kinetics and Mechanism of the Radical Polymerization of Vinyl Monomers
It is known from the theory of frontal polymerization [170–173] that the reaction front velocity is a
power law function of original initiator concentration: u ~ I 0n . The n value obtained by numerical model
ing is 0.40 [170], and that inferred from theoretical analysis is 0.48 [171–173]. Even early experimental
studies [174, 175] showed however that, in the case of the frontal polymerization of 3(oxyethylene)γ,ω
dimethacrylate (OEDMA) and methyl methacrylate at high pressures (up to 5 kbar), n depended on the
nature of the initiator and monomer. In particular, for OEDMA frontal polymerization initiated by di
tertbutyl peroxide (tertBP), benzoyl peroxide (BP), and dicyclohexyl peroxydicarbonate (DCPC), n was
determined to be 0.22, 0.32, and 0.34, respectively. For methyl methacrylate (with BP as an initiator),
n = 0.36. Based on only these results, one might assume that the variation in n is due to a specific effect of
high pressures on the efficiency of initiation, chain breaking, etc. However, in later studies [176] of the
normalpressure frontal polymerization of methacrylic acid in the presence of AIBN, cumyl peroxide
(CP), lauryl peroxide (LP), and ditertbutyl peroxide (tertBP) and that of triethylene glycol dimethacry

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


18 DAVTYAN et al.

T – T0, °C
60
50
40
30
20
10

0 0.2 0.4 0.6 0.8 1.0


α

Fig. 24. T – T0 as a function of conversion [Eq. (8)] for the activated anion polymerization of εcaprolactam; T0 = 150°C,
initial catalyst (C0) and activator (A0) concentrations C0 = A0 = 3.35 × 10–2 mol/L, M0 = 9.543 mol/L.

T – T0, °C
80

3 2
60
1
4

40

20

0 20 40 60
Time, min

Fig. 25. (1–3) Adiabatic heating through the net effect of polymerization and crystallization processes and (4) polymer
ization kinetics. T0 = 110°C, C0 = 3.35 × 10–2 mol/L, A0 = 4 × 10–2 mol/L, M0 = 9.54 mol/L. SiO2 content: (1) 0, (2)
3, and (3) 5 wt % relative to the monomer.

(T – T0)/ αM0
9.0
8.5
8.0
7.5
7.0
6.5
6.0
0 0.2 0.4 0.6
h/α

Fig. 26. Plot representing Eq. (7) for the same polymerization conditions as in Fig. 25.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 19

T–T0, °C

30

3
2
20
1

10

0
20 40 60
Time, min

Fig. 27. Effect of SiO2 nanoadditives on the kinetics of adiabatic poly(εcaprolactam) crystallization. The initial concen
trations are the same as in Fig. 26.

(a) 1.6 μm (b) 1.6 μm

Fig. 28. Electron micrographs of poly(εcaprolactam) (a) containing no additives [188] and (b) containing SiO2 nano
particles [19] (5 wt % relative to the monomer).

late (TEGDM) in the presence of AIBN, BP, and LP, n was determined to be 0.24, 0.25, 0.27, 0.26, 0.2,
0.23, and 0.31, respectively.
Thus, it follows from the above data that the reaction order with respect to the initiator in the frontal
polymerization of vinyl monomers depends rather significantly on the nature of the initiator. A natural
question in this context is how RAF formation influences the reaction order with respect to the initiator
and, accordingly, elementary reactions, such as initiation, chain growth, and chain breaking.
Note, first of all, that there are little or no data on the effect of nanoparticulate additives on the kinetics
or mechanism of the radical polymerization of vinyl monomers. Because of this, we will consider only
reports by Tonoyan et al. [12, 14] and Manukyan et al. [177] concerned with the effect of clay additions
on the reaction order with respect to the initiator in acrylamide (AAm) frontal polymerization.
Figure 22 shows the AAm frontal polymerization velocity [14] as a function of initiator concentration.
The data in Fig. 22 were used to determine the front velocity order with respect to BP concentration,
which was found to be ~0.6 at a filler content of 30%. A similar approach was used to determine the reac
tion order with respect to the initiator (BP and AIBN) as a function of filler content for the polymerization
of AAm–bentonite mixtures [12, 14, 177]. The results are presented in Figs. 23a and 23b, respectively.
These data demonstrate that the reaction order with respect to the initiator is an intricate function of
the mixture composition and the type of the initiator. As follows from Fig. 23, the reaction order is inde
pendent of the reaction front direction (downward or upward propagation). For both initiators, n
increases with increasing filler content. When frontal polymerization is initiated by BP, increasing the
filler content leads to an increase in the reaction order with respect to the initiator up to 0.61 ± 0.02,
whereas with AIBN the reaction order increases from 0.38 to 0.58 ± 0.02. The observed increase in the

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


20 DAVTYAN et al.

Pp, a
175
150
125
100
75
50
25

10 20 30 40 50
Time, min

Fig. 29. Time dependence of the average macromolecule length in an amorphous polymer: (䊉) nucleation; SiO2 content:
(䊊) 0, (䉭) 3, and (䊐) 5 wt % relative to the monomer.

reaction order with respect to the initiator to above 0.5 with increasing filler content can be accounted for
in terms of the interaction between initiator molecules and the surface of individual filler layers. Further
decomposition of the layers, initiation, and polymer chain growth take place on the surface of individual
bentonite layers. It is reasonable to assume that the occlusion of polymerization centers is contributed by
the interaction between binder molecules and the surface of individual bentonite layers, which leads to
RAF formation.
It is of interest to note that, in the case of the frontal polymerization of AAm with diatomite additions,
the following results were obtained: n = 0.65 ± 0.02 for BP and n = 0.6 ± 0.02 for AIBN. The increase in
n in the case of diatomite additions is attributable to the intercalation of polyacrylamide macromolecules
into micro and nanopores of diatomite, followed by the interaction of the macromolecules with the pore
surfaces and occlusion of active centers in the pores of the filler.
Based on the above data, it is reasonable to assume that, in the case of the fillers under consideration,
the initiators decompose both in the bulk of the reaction medium and on the surface of nanoparticles and
the termination of growing radicals follows bimolecular and monomolecular (radical occlusion) mecha
nisms. It is these factors which are responsible for the marked increase in the reaction order with respect
to the initiator in AAm frontal polymerization in the presence of bentonite and diatomite.
The difference between the reaction orders in the initiator for mixtures loaded with bentonite and diat
omite can most likely be understood in terms of the structure of the fillers and the nature of the chemical
compounds forming on the surface layers of the bentonite and diatomite.
Thus, analysis of the influence of various nanofillers on the macrokinetic behavior of the frontal radical
polymerization of vinyl monomers is an effective approach for gaining insight into the kinetics and mech
anisms of the formation of polymer nanocomposites.

2.8. Effect of SiO2 Nanofiller on Crystallization Kinetics under Adiabatic εCaprolactam Anionic
Polymerization Conditions
The influence of nanoadditives on the kinetics of polymerization and those of the crystallization of the
forming polymer has been the subject of extensive studies [178–185]. There is ample evidence [178–191]
that montmorillonite [178, 179], tetraethyl orthosilicate [180], and quartz [181] nanoadditions increase
the cure rate of epoxy compounds. According to Joo Young Nam et al. [181], quartz nanoparticles
increase both the cure rate and the degree of crosslinking.
Similar results were obtained in the synthesis of nanocomposites containing thermoplastic binders. For
example, Lincoln and Vaia [182] showed that, in ethylene polymerization, the addition of clay nanopar
ticles (5 wt %) increased the polymer crystallization rate by three times. The most likely reason for the
increase in crystallization rate is that the clay nanoparticles act as heterogeneous nucleation centers for
the crystallization of the polymer formed. This hypothesis is supported by results of Amerio et al. [183],
who studied the nonisothermal crystallization kinetics of nanocomposites prepared by adding montmo
rillonite to molten polyethylene. Moreover, montmorillonite additions were reported to increase the crys

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 21

tallization rate (during melt cooling) of polylactide [184] and polyamide6 [185] and the polymerization
rate of urethanes [186]. In particular, Rosso and Lin Ye [185] showed that the nonisothermal crystalliza
tion kinetics of polyamide6 were determined not only by nanofiller additions but also by polyamide
6/montmorillonite nanocomposite additions.
Thus, it follows from the above data that the addition of various nanoparticles may influence not only
the kinetics of polymerization but also those of the crystallization of the forming polymer. A natural ques
tion in this context is which process (polymerization or crystallization) can be influenced by a nanopar
ticulate filler if the forming polymer may crystallize in the course of polymerization? This formulation of
the problem is of particular interest when the kinetics of an overall process are studied by thermometric
techniques, e.g., by isothermal or adiabatic calorimetry. Clearly, in such cases the measured total released
heat is contributed by both polymerization and crystallization. There are then serious difficulties in gain
ing insight into the mechanism behind the influence of nanoadditives on these processes.
A typical example of combined polymerization–crystallization processes is the activated anionic poly
merization of lactams, which is of practical interest for the fabrication of various articles by “chemical
forming” processes. As shown by Benson and Cairos [187], the activated anionic polymerization of lac
tams is accompanied by selfheating of the reaction mixture because both the polymerization and the
crystallization of the forming polymer are exothermic processes. Even though the heat effect of the poly
merization process is rather small, e.g., just 13.4 kJ/mol for εcaprolactam, heating during the polymer
ization process may reach 90–92°C. The polymerization is accompanied by an exothermic crystallization
of the forming poly(εcaprolactam), with a heat effect of 25 kJ/mol, which contributes to the reaction
temperature. It is known [188–190] that, at initial process temperatures below 150°C, the polymerization
of εcaprolactam and the crystallization of the forming polymer occur simultaneously, whereas above
150°C poly(εcaprolactam) crystallizes only after completion of the polymerization process. When poly
merization and crystallization processes take place simultaneously, there are serious difficulties in gaining
insight into the mechanism behind the influence of nanoadditives on these processes. In view of this, con
sider, like in a previous study [191], the possibility of separately examining parallel polymerization and
crystallization processes and assessing the influence of SiO2 nanoadditions on the kinetics of adiabatic
anionic polymerization of εcaprolactam.
Under the conditions of the adiabatic activated anionic polymerization of εcaprolactam, the rate of
the overall increase in reaction temperature is contributed by the polymerization and crystallization pro
cesses [188, 189]:
dn
cρ dT = Qp dn + Qcr cr , (4)
dt dt dt

where c and ρ are the specific heat and density of the reaction mixture; Т is the temperature of the reaction
medium; n and ncr are the amount of the polymer formed and that of the crystallized material; Qp and Qcr
are the heats of the polymerization and crystallization, respectively; and t is time.
The initial conditions for Eq. (4) have the form t = 0, T = T0, and n = ncr = 0.
Assuming that the Qp /cρ and Qcr/cρ are independent of temperature and conversion [188, 189], we
obtain by integrating Eq. (4)
Q Q
T − T0 = p n + cr ncr , (5)
cρ cρ

where T0 is the initial reaction temperature.


Conversion (α ) and the degree of crystallinity (η) can be expressed through n and ncr using the relations
M −M n
α= 0 = n ; β = cr . (6)
M0 M0 M0

Here, M0 and M are the initial and instantaneous monomer concentrations. Substituting (6) in (5), we
obtain
i i
Q Q
T − T0 = p α M 0 + cr β M 0. (7)
cρ cρ
In relation (7), T = Tp + Tcr, where Tp is the adiabatic heating through polymerization, and Tcr is that
through crystallization. When the polymerization and crystallization processes are separated in time
(Т0 ≥ 150°С), we have η = 0 during polymerization. Therefore,

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


22 DAVTYAN et al.

Table 1. SC properties of polymer–ceramic nanocomposites based on Y1Ba2Cu3O6.97 (tpr = 200°C, τpr = 30 min)
Ceramic : binder
Composition Tc, K Tf, K Note
weight ratio
80 : 20 96 84
UHMWPE + Y1Ba2Cu3O6.97 85 : 15 96 84 Prepared after reduction
85 : 15 96 84
BPE + Y1Ba2Cu3O6.97 80 : 20 94 80 ''
ETFE + Y1Ba2Cu3O6.97 75 : 25 96 77 ''
BP + Y1Ba2Cu3O6.97 80 : 20 96 83 ''
PVDF + Y1Ba2Cu3O6.97 85 : 15 90 75 ''
PFA + Y1Ba2Cu3O6.97 80 : 20 88 76 ''
UHMWPE + irganox + Y1Ba2Cu3O6.97 80 : 20 96 89 Prepared with no reduction
BPE + irganox + Y1Ba2Cu3O6.97 80 : 20 94 85 ''
PVA + irganox + Y1Ba2Cu3O6.97 85 : 15 90 80 ''

Qp M 0
T = Tp − T0 = α. (8)

Preliminary experiments at Т0 ≈ 155°С, where εcaprolactam polymerization and the crystallization of
the forming poly(εcaprolactam) were separated in time, showed that relation (8) was satisfied (Fig. 24).
Substituting the measured slope in (8), we obtain
Tp − T0 = 52α . (9)
From (9), we find that the heat of polymerization is Qp = 13.6 kJ/mol.
On the other hand, the straight line in the plot representing Eq. (8) indicates that Qp/cρ is constant
when polymerization and crystallization are separated in time. Under the assumption that Qp/cρ remains
constant as well when polymerization and crystallization take place simultaneously, relation (9) can be
used to find T – Т0 from α.
Curves 1–3 in Fig. 25 are net kinetic curves showing the increase in temperature due to polymerization
and crystallization.
Curve 4 in Fig. 25 represents the kinetics of the polymerizationinduced heating inferred from relation
(9) using gravimetric data obtained at various SiO2 contents. From the difference between curves 1–3 and
4 in Fig. 25, we obtained kinetic curves for adiabatic crystallization (Fig. 25, curve 3).

χ
1.0

0.8

0.6

0.4

0.2

0 50 100 150
Temperature, K
Fig. 30. SC transition of a UHMWPE + Y1Ba2Cu3O7 – x polymer–ceramic nanocomposite. Ceramic : binder weight
ratio, 80 : 20; pressing temperature, 200°C.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 23

Table 2. SC properties of nanocomposites based on Y1Ba2Cu3O6.92 ceramic (tpr = 200°C, τpr = 30 min)
Antioxidant
Polymer binder in the
Wt % binder (5 wt % relative to Tc, К ΔT, К
nanocomposite
the binder)
HPPE 10 Irganox 92.1 7.0
15 NG2246 91.8 7.0
10 91.2 7.0
15 91.5 ∼5.0
UHMWPE 20 91.8 6.0
15 Irganox 91.7 6.0
15 NG2246 91.2 6.0
10 92.3 8.0
15 93.7 7.0
PMMA
20 91.7 7.0
15 NG2246 93.2 6.0
10 92.0 6.0
15 93.1 7.0
PS
20 92.3 8.0
15 NG2246 93.0 7.0
10 91.7 7.0
Poly(StcoMMA) 15 92.3 7.0
(60 : 40 mol %) 20 92.1 7.0
15 NG2246 92.0 7.0
10 92.1 8.0
Poly(StcoMMA) 15 91.9 7.0
(80 : 20 mol %) 20 92.6 8.0
15 NG2246 93.4 ∼9.0

From relation (7) and the kinetic curves in Fig. 25, we find that the plot of (T – T0)/αM0 versus η/α
has the form of a straight line (Fig. 26). The values of α and η can then be found as
T −T Tcr − T0
α = p∞ 0 , η = , (10)
Tp − T0 Qp ∞
(Tp − T0 )
Qcr
where Tp and Tcr are timedependent polymerization and crystallization temperatures.
The linear plot in Fig. 25 validates the assumption that Qp/cρ is constant when polymerization and
crystallization processes take place simultaneously. Note that the Qp values determined from Figs. 23 and
25 are identical. The heat of crystallization evaluated from Fig. 26 is Qcr = 25.6 kJ/mol.
It follows from curves 1–3 in Fig. 25 that, with increasing SiO2 content, the reaction temperature rises
more rapidly, i.e., the rate of adiabatic polymerization increases.
It is of interest to note that the kinetic curves of the adiabatic increase in reaction temperature diverge
only 13–15 min after the polymerization onset. The likely reason for this is that SiO2 nanoparticles
increase the crystallization rate of the forming polymer, without influencing the polymerization kinetics.
Indeed, it is seen in Fig. 27 that the addition of SiO2 nanopowder influences the crystallization rate
throughout the process. The highest adiabatic crystallization temperature is independent of the percent
age of added SiO2. The data presented in Figs. 27 and 28 can be accounted for by the fact that nanoparti
cles act as heterogeneous nucleation centers for the crystallization process. Because of this, an increase in
nanofiller content accelerates the entire crystallization process.
To validate the above inferences, samples containing 5% SiO2 nanopowder and those with no filler were
examined by electron microscopy in the initial stages of crystallization. The results are illustrated in Fig. 28.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


24 DAVTYAN et al.

(a) (d)

005 005
113 113

(b) (e)

(c) (f)

Fig. 31. Equilibrium states of phases and deviations from equilibrium as functions of PMMA content (a–c) before and
(d) after the superconducting properties were restored. Compositions of the YBa2Cu3O6.92 : PMMA composites: (a) 93 :
7, (b) 90 : 10, (c) 84 : 16, (d) 95 : 5, (e) 93 : 7, (f) 90 : 10.

Comparison of Figs. 28a and 28b indicates that, in the initial stages of the process, the number of nuclei
is considerably greater in the presence of SiO2 nanoparticles (Fig. 28b), which lends support to the
assumption that SiO2 nanoparticles lead to heterogeneous nucleation in the crystallization process.
It is worth noting that the time delay of the crystallization onset (Fig. 25, curves 1–3) relative to the
polymerization (Fig. 25, curve 4) is caused by the formation of poly(εcaprolactam) with a molecular
weight necessary for its crystallization.
Comparison of the crystallization onset time (Fig. 27, curves 1–3) and polymerization kinetics (Fig. 2,
curve 4) allows us to estimate the critical average poly(εcaprolactam) molecule length at which the crys
tallization process begins. It follows from Figs. 25 and 27 that the degree of polymerization corresponding
to the crystallization onset is approximately 0.14. Therefore, assuming that all of the activator had been
consumed by the crystallization onset, we obtain for the critical average macromolecule length (Pp) cor
responding to the crystallization onset
Pp = α i ⋅ M 0 A0 , (11)
where αi is the degree of conversion corresponding to the crystallization onset, and M0 and A0 are the ini
tial monomer and activator concentrations, respectively. An estimate using Eq. (11) indicates that the
average macromolecule length at which crystalline nuclei appear is Pi = 33.
In addition, using the data in Figs. 25 and 27 we can estimate the critical average length of growing mac
romolecules (that are entirely in the amorphous phase of the polymer) involved in the formation of crystal
line polymer. Clearly, the fraction of amorphous polymer in the polymerization process is (α – η)M0, where
η is the degree of crystallinity. The average macromolecule length (Pp, a) in the amorphous polymer is then
Pp, a = (α – β)M0/A0, (12)
where α and η are defined by (10).
Comparison of the experimental data presented in Fig. 25 (curves 1–4) and 27 (curves 1–3) with rela
tions (10) and (12) allows us to determine Pp, a as a function of time during polymerization. Qp and Qcr were
taken from Refs. [188–190].
Figure 29 plots Pp, a versus time at various SiO2 nanopowder contents.
As seen in Fig. 29, the time variation of Pp, a is essentially independent of SiO2 content. As pointed out
above, the reason for this is that the nanoparticles are involved in the crystallization process only as heter
ogeneous nucleation centers. The increase in the slope of the plot of Pp, versus time is due to the adiabatic
increase in reaction temperature, which has a stronger influence on the polymerization rate than on the
crystallization process.
Thus, studies of parallel polymerization and crystallization processes, analysis of experimental data,
and SEM examination of the nanocomposites obtained demonstrate that the crystallization kinetics of

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 25

Table 3. Thermal stability of polymer binders and related superconducting polymer–ceramic nanocomposites (tpr =
130°C, τpr = 4 min)
Antioxidant, Thermooxidative Weight loss at
Binder Wt % binder 5 wt % relative destruction onset 300°C, % relative Note
to the binder temperature, °C to the binder
UHMWPE 100 195 2.3 Ceramic with reduced
10 185 ∼33.0 oxygen content
15 190 ∼21.0
20 195 ∼20.0
PMMA 100 NG2246 170 21.0 Standard Y1Ba2Cu3O6.97
15 155 33.0 ceramic
15 235 21.0
Poly(StcoMMA) 100 NG2246 165 20 ''
(60 : 40 mol %) 15 125 ''
15 220 ∼5 ''

poly(εcaprolactam) are determined by SiO2 content, because silica particles participate in the nucleation
process.

2.9. Thermodynamic Aspects of Nanoparticle–Macromolecule Interaction


Analysis of the literature [12–14, 67–70, 114, 139, 147, 148] and comparison with theoretical concepts
[192–198] suggest that there is a qualitative correlation. Consider briefly the relevant results.
Nanoparticles synthesized in the presence of macromolecules are known to have a rather narrow size
distribution [192–198]. The influence of the nature of macromolecules on the size distribution, agglom
eration behavior, and stability of nanoparticles depends rather strongly on synthesis conditions, in partic
ular on the process temperature, the thermodynamic properties of the solvent, and other factors.
As shown by Litmanovich et al. [193], the average particle sizes of nickel and copper recovered from
appropriate aqueous salt solutions in the presence of poly(Nvinylpyrrolidone) increase with temperature.
At the same time, the average size of copper nanoparticles decreases with increasing temperature if the
4 copper is reduced by ions from aqueous poly(Nvinylcapropropyllactam) [194]. Moreover, Litmanovich
et al. [195] showed that the average size, size distribution, and agglomeration behavior of nanoparticles
depended rather strongly on the molecular weight and polydispersity index of the polymer.

Weight loss, %
Weight loss, % 1
100
1
100

75 90

50
2 80 2
25

0 100 200 300 0 100 200 300


Temperature, °C Temperature, °C

Fig. 32. (1) Weight loss (percent of the initial weight of Fig. 33. (1) Weight loss (percent of the initial weight of
the polymer) and (2) DTA curve for the thermooxida the binder) and (2) DTA curve for a nanocomposite
tive destruction of UHMWPE at a heating rate of containing UHMWPE as a binder. UHMWPE :
3.2 K/min. Y1Ba2Cu3O6.97 = 20 : 80 (wt %); heating rate, 3.2 K/min.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


26 DAVTYAN et al.

χ
1.0

0.8

0.6

0.4

0.2

0
80 90 100
Temperature, K
Fig. 34. SC transition of a nanocomposite produced by polymerizing MMA in pressed Y1Ba2Cu3O6.97 ceramic. Ceramic
: binder weight ratio, 88 : 12.

To interpret these results, Litmanovich et al. [192–198] assume that the adsorption of polymer mac
romolecules on the surface of metallic nanoparticles can be described by an equilibrium constant (K) and
the numberaverage molecular weight of the binder. A high molecular weight polymer is thought to be
capable of adsorbing larger nanoparticles [192]; that is, a long macromolecule can interact with more than
one particle. Therefore,
K = K 1x ,
where K1 = exp(–ΔG1/RT), ΔG1 is the average integral energy of bond formation between the macromol
ecule and nanoparticle surface, and x is proportional to the surface area of the growing particles.
Clearly, noncovalent interaction between polymer macromolecules and the surface of nanoparticles
may have a van der Waals, electrostatic, hydrogen, or another nature. If interaction is not covalent but has
any other nature, ΔG1 is, as a rule, rather low. On the other hand, since low x values correspond to a small
equilibrium constant, the formation of a stable complex between a macromolecule and large particles is
in principle impossible.
Conceptually, nanoparticle growth is accompanied by a rather rapid increase in K. At a certain critical
size of nanoparticles, their entire surface becomes covered with polymer macromolecules. Subsequently,
the nanoparticles do not grow. This process was analyzed quantitatively by Gole and Murphy [199].
According to their results, the size distribution n(r) of spherical nanoparticles has the form
⎛ r ⎞

r2 −α −α r 2
K 1 ) exp ⎜ − ln(1 + c0 N K 1 )dr ⎟ ,
n(r ) = ln(1 + c0 N
⎜ ⎟
⎝ 0 ⎠
where с0 is the initial monomer concentration and N is the numberaverage molecular weight. For good
solvents, ν = N0.8; for θsolvents, ν = N0.5. The accuracy of the equation derived by Gole and Murphy
[199] is determined by the high affinity of macromolecules for nanoparticles.
According to Gole and Murphy [199], the average nanoparticle size is given by
α ln N − ln c0
r2 =
ln K 1
It follows from this relation that, from the average nanoparticle size as a function of numberaverage
molecular weight, one can evaluate α, K1, and, therefore, ΔG1. In addition, knowing the average nanopar
ticle size as a function of temperature, one can evaluate ΔH and ΔS.
The results reported by Litmanovich et al. [193] lead us to conclude that, even when the energy of the
interaction between macromolecules and growing particles is comparable to thermal energy, kT, the form
ing stable nanoparticles should have a polymer coating on the order of 1 nm in thickness and a narrow size
distribution. Thus, we conclude from the above data that the thermodynamic model considered in this
section is an important tool for the synthesis of nanoparticles and polymer nanocomposites and also for
gaining insight into the mechanisms of polymer–nanoparticle interactions.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 27

Q, cal/g−1
1
120 2
3

80
4

40

0 200 400 600 Time, min

Fig. 35. Effect of Y1Ba2Cu3O6.97 content on the polymerization kinetics of MMA at 60°C in the presence of 3 × 10–3 mol/L
of AIBN. The amount of the ceramic is (1) 0, (2) 0.1, (3) 0.2, and (0.4) 0.4 g.

Q, cal/g−1 (a) Q, (b)

8
120 120
7
6
80 5 80
4
3
2
40 1 40

0 200 400 0 1 2
Time, min Y1Ba2Cu3O7 − x, g
Fig. 36. (a) Polymerization kinetics of MMA in the presence of Y1Ba2Cu3O6.97 ceramics at 60°C. The amount of the
ceramic is (1) 0, (2) 0.1, (3) 0.2, (4) 0.3, (5) 0.5, (6) 0.7, (7) 1, and (8) 3.8 g. (b) Q as a function of the amount of the
ceramic.

3. SUPERCONDUCTING INTERCALATED POLYMER–CERAMIC NANOCOMPOSITES


3.1. Superconducting Properties of Polymer–Ceramic Nanocomposites Prepared by Hot Pressing
Studies of the effect of the nature of polymer binders on the superconducting properties of polymer–
ceramic nanocomposites produced by pressing at 200°С for 30 min (with an arbitrary cooling schedule of
the die) showed [100–104, 107] that the materials thus obtained had no superconducting properties (no
Meissner signal was detected). The lack of superconducting properties may be the result of oxygen removal
5 from the superconducting orthorhombic phase of the ceramic during the forming process. The released
oxygen may be involved in the thermooxidative destruction of the polymer matrix. It seems, however,
unlikely that this is the oxygen that desorbs from the grains of the ceramic and diffuses into the polymer.
It seems likely that it is the free oxygen located on the surface of the oxide ceramic grains that reacts with
the polymer. This assumption is indirectly supported by results demonstrating that the superconducting
properties of nanocomposites can only be restored by exposure to a dry oxygen flow at temperatures of the
αtransition in polymer binders (Table 1).
More conclusive evidence in favor of the above assumption is provided by results of experiments in which
an antioxidant was added to a starting mixture for the preparation of a nanocomposite [101, 103, 107].

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


28 DAVTYAN et al.

Antioxidant additions to a polymer matrix are known to markedly reduce the oxidative destruction rate
in polymers [200, 201].
It is therefore reasonable to expect that the addition of small amounts (0.5 wt % relative to the binder)
of an antioxidant to a polymer–ceramic nanocomposite will reduce the oxidation rate, thereby inhibiting
the oxygen removal from the surface of the ceramic grains.
As seen from Table 1, nanocomposites with a low antioxidant content have superconducting properties
just after the forming step, with no reduction.
The restoration of the superconducting properties of polymer–ceramic nanocomposites in a dry oxy
gen flow increases their superconducting onset temperature (Tc) by 2–3 K and changes their zeroresis
tance temperature. The broadening of the transition is most likely caused by a nonuniform distribution of
the oxygen over the surface of the ceramic grains after the reduction step.
Figure 30 shows a typical superconducting transition curve for a polymer–ceramic nanocomposites
after the reduction step.
Comparison of the superconducting properties of polymer–ceramic nanocomposites based on various
binders (Table 1) that both the onset and end point of the transition depend significantly on the nature and
chemical composition of the polymer binder.
It can be seen from the data in Table 1 that the superconducting transition temperature of the super
conducting polymer–ceramic nanocomposites is 23 K higher, except for the materials based on PVA and
polyformaldehyde. The increase in Tc can be tentatively attributed to physical interaction of particular
groups of binder macromolecules with the ceramic surface, to the point of intercalation [108–112] of such
groups into the interlayer spaces of the ceramic grains or “anchoring” of them on vacant oxygen sites.
The assumed interaction of individual chemical elements (or groups) in polymer chains with the ceramic
surface may be evidenced by the decrease in Tc by 5 K in the case of a nanocomposite containing PVA as
a binder. Like water [202], alcohols [203], or weak acids [204], the OH groups of this binder enter into
5 chemical reactions and distort the orthorhombic phase in the surface layer of the ceramic grains, thereby
reducing the superconducting transition temperature.
The large decrease in Tc in the case of the polyformaldehydematrix nanocomposite can be accounted
for by the increased tendency of polyformaldehyde toward oxidative thermal destruction, which leads to
increased oxygen removal from the ceramic. Moreover, even at relatively low temperatures [205] the ther
mal destruction of polyformaldehyde is accompanied by depolymerization and gaseous formaldehyde
evolution, which may have an adverse effect on the superconducting properties of the nanocomposite.
At the same time, more rigorous validation of the above assumptions required additional, more detailed
work [108–112], which addressed the influence of both the nature of polymer binders and preparation
conditions on the superconducting properties of the materials.

3.2. Effects of Thermal Conditions and Antioxidant Additives on ThermoOxidative Destruction Processes
To ensure the tightest and strongest contact between a polymer binder and ceramic grains, starting
mixtures should be pressed at temperatures slightly above the melting point or glass transition temperature
of the polymer matrix. The glass transition temperatures of the acryl and vinyl homo and copolymers
commonly used for such purposes do not exceed 100°С, and the melting points of polyolefins lie in the
range 120–125°С (except for polypropylene, which has tm = 165–170°С). In view of this, most samples
were prepared by pressing at 130°С. In contrast to the above case, the SC polymer–ceramic nanocom
posites produced by pressing at 130°С retain their SC properties. Note that polyethylene binders ensure
roughly the same Tc and ΔТ values as in the parent ceramics, and nanocomposites containing acryl and
vinyl homo and copolymers have 1 to 2°С higher SC onset temperatures (Table 2). There is no definite
correlation between the ceramic content of SC polymer–ceramic nanocomposites and their Tc and ΔТ in
the range of filler contents studied (80–90 wt %). Raising the compaction pressure to 200 MPa has no
effect on Tc or the transition width (the nanocomposite containing 15 wt % UHMWPE has Tc = 92 K and
ΔТ ≈ 8 K).
As pointed out above, pressing polymer–ceramic nanocomposites at 200°С for 30 min completely
degrades their SC properties. Reducing the pressing time to 4 min, without changing the temperature,
allows UHMWPEbased nanocomposites (85 wt % ceramic) to have the same SC transition parameters
as in the parent ceramic (Tc = 91.6 K, ΔТ ≈ 6.5 K).
A different picture is observed when PPbased nanocomposites of the same composition are produced
at 180°С: the superconducting transition temperature remains unchanged at that of the parent ceramic
(91.6 K), whereas the transition width increases sharply (ΔТ ≥ 8 K).

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 29

Table 4. Superconducting and physicomechanical properties of Mn, Co, Zn, and Nicontaining polymer–ce
ramic nanocomposites
Composition
metal complex Pressing σ, E ×10–3
Y1Ba2Cu3O6.98 Metal Tc, К ΔT, К ε, %
monomer time, min MPa MPa
g wt % g wt %
0.293 43 0.388 57 Mn 10 95 8 – – –
0.396 50 0.396 50 Mn 5 94 9 – – –
0.90 70 0.290 30 Mn 5 94 10 15 4.55 3
0.518 73 0.196 27 Mn 5 93 8 – – –
0.416 67 0.209 33 Mn 10 94 11 – – –
0.552 78 0.156 22 Mn 5 95 10 – – –
0.325 51 0.318 49 Co 5 93 10 – – –
0.432 60 0.283 40 Co 5 92 8 – 2.40 –
0.503 70 0.228 30 Co 2 92 8 17 4.2 5
0.90 70 0.390 30 Zn 5 95 10 15 – 4
0.416 67 0.209 33 Zn 10 94 11 – 3.10 –
0.552 78 0.156 22 Zn 5 95 10 – – –
0.486 70 0.208 30 Ni 5 95 8 14 – 2

It is known that the presence of tertiary carbon atoms in PP macromolecules leads to weakening of the
carbon–carbon bonds. As a result, PP is less stable than PE [206]. It seems likely that it is the increased
tendency of PP toward thermooxidative destruction and the participation of oxygen from the supercon
5 ducting orthorhombic phase in this process which are mainly responsible for the observed fact [94, 101,
111, 112].
To validate this assumption, antioxidants (Irganox and NG2246) were introduced into a matrix in
order to reduce the rate of thermooxidative destruction of the polymer. Indeed, Irganox additions were
found to reduce the superconducting transition width of the nanocomposite to the level of the parent
ceramic (ΔТ ≈ 6.5 K). Note that NG2246 has no effect on the ΔТ of the PPbased materials. The most
likely reason for this is that NG2246 is a rather poor antioxidant for polyolefins. Raising the pressing tem
perature of a PMMAbased nanocomposite (85 wt % ceramic) to 160°С or above (pressing time, 15 min;
arbitrary cooling schedule of the die to 40°С) considerably increases the ΔТ of the resultant materials
(≥8 K), with Tc remaining in the range 91.9–92.3 K.
5 In connection with this, the equilibrium state of the tetragonal and orthorhombic phases in polymer–
ceramic nanocomposites containing PMMA as a binder and possible deviations from equilibrium were
5 studied by Xray diffraction [94]. The equilibrium state and deviations to the tetragonal or orthorhombic
phase were assessed from variations in the relative intensity of the 005 (014) and 113 reflections in the
range 2θ = 38–41°, sensitive to oxygen content [207]. It is known that, at δ ≈ 1 (where δ is the oxygen
content of the nanocomposites), the 113 reflection is stronger than the 005 (014) reflection, and the ratio
of their intensities is 1.35 [208], whereas the relative intensity of the strongest line, 110 (103), is insensitive
to oxygen content. Figure 31 illustrates the equilibrium states of phases and deviations from equilibrium
as functions of PMMA content for a polymer–ceramic nanocomposite containing PMMA as a binder.
Based on Xray diffraction data, Arakelova et al. [94] concluded that the phases present in the nano
composites were in the equilibrium state at ceramic : polymer ratios of 93 : 7 and 90 : 10 (Figs. 31a, 31b).
The 84 : 16 composite had an orthorhombically distorted lattice (Fig. 31 c).
Of special note is that these results are characteristic of the nanocomposites produced by pressing at
200°С for 30 min, with no reduction step. It is of interest to note that, after restoring the superconducting prop
5 erties in a dry oxygen flow [108–112], orthorhombic distortion was observed at ceramic : polymer = 93 : 7 and
90 : 10 (Figs. 31e, 31f), whereas the equilibrium state of the phases was observed in the 95 : 5 nanocom
posite (Fig. 31d). It is worth pointing out that the addition of the NG2246 antioxidant to this nanocom
posite or reducing the pressing time to a standard 4 min (under a standard die cooling schedule) restored
its SC properties (Tc = 91.5 K, ΔТ ≈ 6 K).

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


30 DAVTYAN et al.

Table 5. Physicomechanical properties of superconducting polymer–ceramic nanocomposites


Ceramic : binder Tempera σ, E,
Composition Deformation ε, %
weight ratio ture, K MPa MPa
UHMWPE + Y1Ba2Cu3O6.92 80 : 20 Tension 300 30 100 10
85 : 15 77 100 – 0.1
HPPE + Y1Ba2Cu3O6.92 80 : 20 – 300 15 75–80 9–10
ETFE + Y1Ba2Cu3O6.92 75 : 25 – 300 32 150 7.2
BP + Y1Ba2Cu3O6.92 80 : 20 – 300 28 100–110 8.3
PVA + @@@ + Y1Ba2Cu3O6.92 85 : 20 – 300 34 130 7.5

3.3. Effect of ThermoOxidative Destruction on the Superconducting Properties


of Polymer–Ceramic Nanocomposites
Thermal analysis data for YBa2Cu3O6.97 ceramics (Table 3; Figs. 32, 33) demonstrate that the SC
ceramics are stable up to 300°С: their thermogravimetric (TG) and differential thermal analysis (DTA)
curves have no breaks or extrema.
Analysis of the DTA data for unfilled UHMWPE (Fig. 32) and a UHMWPEbased nanocomposite
(Fig. 33) leads us to conclude [100–103, 106, 111] that the exothermic peaks in their DTA curves in the
range 150–195°С are most likely due to the oxidation of the matrix at these temperatures.
This is also evidenced by the marked weight gain seen in their TG curves (Figs. 32, 33, curves 1). The
oxidation onset temperature of the UHMWPEbased nanocomposites is 10–15°С lower (Fig. 33) than
that of the unfilled polymer. As seen in Figs. 32 and 33, the thermooxidative destruction of the nanocom
posite begins in the range 185–195°С and is accompanied by a weight loss.
This means that the formation of the UHMWPEcontaining nanocomposites at 200°С is accompa
nied by the thermooxidative destruction of the binder. The PMMA and PSbased polymer–ceramic
nanocomposites containing Y1Ba2Cu3O6.97 ceramics have lower thermal stability. The introduction of the
SC ceramic into a polymer matrix (85 wt % filler) reduces the thermooxidative destruction onset tem
perature of PMMA from 170 to 155°С. At the same time, the addition of the NG2246 antioxidant to such
a nanocomposite considerably reduces the rate of thermooxidative processes in the matrix, thereby rais
ing the decomposition onset temperature of the binder to 235°С. These data are thought to completely
validate the above assumptions [102–104] and are quite consistent with the general trends observed in
studies of the SC properties of polymer–ceramic nanocomposites containing PMMA and PS as binders.
The NG2246 antioxidant is equally effective for the nanocomposite based on a styrene–methyl meth
acrylate copolymer as a binder (Table 3) and for that containing poly(methyl methacrylate) as a binder
(Table 3).

σ, kgf/mm2 (a) σ, kgf/mm2 (b)


3.00
10.0

2.25
7.5

1.50 5.0

0.75 2.5

0 5 10 0 0.1
ε, % ε, %
Fig. 37. Stress–strain diagrams of nanocomposites containing UHMWPE as a binder. Composition of the nanocompos
ite: Y1Ba2Cu3O6.97 : UHMWPE = 90 : 10. Test temperature: (a) 300 and (b) 77 K.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 31

Table 6. Effects of particle size and filler content on σ, E, ε, Tc, and Tf


Average size, Sp. surface,
Wt % filler σ, MPa E, MPa ε, % Тc, °С Тf, °С
μm cm2/g
10 18.5 65 280
20 26.0 85 260
5 1132 30 55.0 115 210
40 100.0 151 160
50 175.0 190 100 87 73
10 11.0 55 277
20 20.0 74 263
15 755 30 45.0 100 230
40 90.0 132 183
50 170.0 165 130 91 83
10 8.0 50 275
20 18.0 68 264
25 453 30 40.0 90 225
40 72.0 120 200
50 155.0 150 157 95 88
10 8.0 48 272
20 16.0 60 270
35 323 30 35.0 80 252
40 620.0 100 120
50 155.0 130 180 95 89

Thus, all relevant experimental data (Table 3) can be accounted for in terms of the competition of two
concurrent processes:
(1) interaction of individual elements or groups in binder macromolecules with the surface of ceramic
grains to the point of intercalation of such groups into the interlayer spaces of the ceramic and
(2) thermooxidative destruction of the polymer matrix.
Factors that both contribute to the interaction of elements of binder macromolecules with the ceramic
surface (flexibility and composition of the macrochains, heating, and others) and reduce the thermooxi
dative destruction rate (decrease in pressing temperature or addition of an antioxidant) enable the the
superconducting transition parameters of the ceramic in the nanocomposites to be improved or, at least,
retained.
The above data have a number of important implications.
The pressing of SC polymer–ceramic nanocomposites at temperatures corresponding to a thermally
stable state of the polymer binder typically ensures SC transition parameters similar to those of the parent
ceramic. The slightly (1–3°С) higher Tc of the nanocomposites containing acryl and vinyl homo and
copolymers as binders is due to the interaction of particular groups of binder macromolecules with the
ceramic surface and, as mentioned above, the intercalation of such groups into the interlayer spaces of the
ceramic.
The broadening of the SC transition, i.e., the larger ΔT of the SC polymer–ceramic nanocomposites
produced at elevated temperatures (160°С or above) is caused by the thermooxidative destruction of the
polymer binder. Since the decomposition onset temperature of the binder depends on its chemical struc
ture, ΔT most likely depends on the nature of SC nanocomposites. It may be that the SC ceramic itself
contributes to the thermooxidative destruction of the polymer binder, possibly through the participation
of oxygen from the ceramic in these processes. As pointed out above, however, it seems more likely that
the irreversible thermooxidative destruction of the polymer binder involves free oxygen located on the

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


32 DAVTYAN et al.

E ''
2
5
4
1
3
2
1

−150 −100 −50 0 50 100 150


Temperature, °C

Fig. 38. Dynamic elastic modulus as a function of temperature. Ceramic content: (1) 0 and (2) 15 wt %.

tanδ
0.3

0.2

2 0.1

1
–150 –100 –50 0 50 100 150
Temperature, °C
Fig. 39. Mechanical loss tangent as a function of temperature. Filler content: (1) 0 and (2) 15 wt %.

surface of the oxide ceramic grains. This assumption is indirectly supported by the fact that the difference
between the unitcell parameters of the ceramic is either constant or varies only slightly with pressing tem
perature (in the range 130 to 200°С).
The influence of antioxidants is probably related to the considerable drop in the rate of oxidation pro
cesses in the polymer binder and, accordingly, the slower removal of oxygen from the surface layer of the
ceramic grains.

3.4. Methyl Methacrylate Polymerization in the Presence of Ceramic Y1Ba2Cu3O7 – x


SC property measurements for nanocomposites obtained by polymerizing methyl methacrylate
(MMA) in pressed Y1Ba2Cu3O6.97 ceramics [100–103, 106, 107] showed that all of the materials exhibited
a Meissner effect with no subsequent restoration stage. Figure 34 shows the SC transition of nanocompos
ites containing PMMA thus prepared.
Unexpected results were obtained in studies of the influence of Y1Ba2Cu3O6.97 ceramics on the poly
merization kinetics of MMA in the presence of azoisobutyronitrile (AIBN) and with no initiator. At high
filler contents (Y1Ba2Cu3O6.97 : MMA = 90 : 10 and 85 : 15 wt %) in the presence of AIBN, no polymer
ization occurred for a long time (on the order of 4–5 h), whereas with no initiator polymerization reached
completion rather rapidly.
To identify the origin of such anomalies, special experiments were carried out to assess the effect of
Y1Ba2Cu3O6.97 content on the polymerization kinetics of MMA with and without AIBN. Figures 35 and
36a show kinetic curves illustrating the effect of Y1Ba2Cu3O7 – х 6.97 ceramics on the polymerization kinet
ics of MMA with and without AIBN, respectively.
It is seen from curve 1 in Fig. 36 that, with no ceramic, MMA polymerization has usual kinetics. The
heat release reaches about 130 cal/g, which corresponds to the heat of MMA polymerization.
Y1Ba2Cu3O6.97 ceramic additions reduce not only the polymerization rate but also the limiting degree of
polymerization (Fig. 36, curves 2–4).

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 33

Table 7. Dynamic mechanical characteristics of Y1Ba2Cu3O6.97–UHMWPE nanocomposites


tanδ
UHMWPE : Y1Ba2Cu3O6.97 Е', Е', E', Тg, tanδ Тα,
Т, °С (2nd tran tgδ (α)
weight ratio T = 150°С T = 100°С Т = 25°С °С (gl) °С
sition)
100 : 0 3.0 1.5 1.1 –120 0.01 – – 144 0.2
85 : 15 5.1 3.4 1.6 –99 0.06 30 0.015 151 0.25
50 : 50 10.1 6.5 3.1 –94 0.065 30 0.025 155 –
15 : 85 – – 4.5 – – – – 157 0.2

2 Table 8. Effect of filler content on the melting point and enthalpy of fusion of the binder in UHMWPE + Y1Ba2Cu3O6.97
nanocomposites
UHMWPE : Y1Ba2Cu3O6.97
Tm (onset) ΔHm, J/(g UHMWPE) Percent crystallinity
weight ratio
100 : 0 140 115.0 39.1
85 : 15 149 116.5 39.7
50 : 50 137 122.5 41.7
15 : 85 136 123.5 42.0

The opposite is observed when MMA is polymerized with no initiator. As seen in Fig. 36a, this
increases the polymerization rate, with rather unusual transformation kinetics. Ceramic additions lead to
a large increase in polymerization rate in the initial stages of the process and reduce the magnitude of the
gel effect in the latter stages of the transformation. These anomalies become more pronounced with
increasing ceramic content in the starting reaction mixture.
The observed kinetics of MMA polymerization can be understood in terms of specific interactions of
the ceramic with both the monomer and initiator. It is known [209–212] that perovskites, including
Y1Ba2Cu3O7 – x , are capable of catalyzing many chemical reactions. It is therefore conceivable that local
areas on the surface of the ceramic grains may react with MMA and AIBN. If the rate of the reaction of
the ceramic with the initiator is higher than that with the monomer, some of the initiator, reacting with
active areas on the surface of the ceramic (Y1Ba2Cu3O6.97) grains, will completely cover them, preventing
reaction with MMA. Some of the initiator is thus “blocked” by the ceramic surface, without participating
in the initiation reaction. Because of this, increasing the amount of added ceramic increases the percent
age of the “blocked” initiator and, hence, reduces the polymerization rate (Fig. 36a, curves 2–4).
As to the polymerization kinetics in the absence of AIBN, it seems likely that the reaction of the mono
mer with active local areas on the surface of the ceramic grains leads to the formation of primary polymer
ization centers, which remain attached [213] to the solid Y1Ba2Cu3O6.97 surface. In this case, it is obviously
difficult to employ classical concepts pertaining to the mechanisms of chain growth and termination in
the radical polymerization of vinyl monomers because the immobility of the reactive ends of the macro
molecules rules out bimolecular chain termination reactions. This can account for the large increase in
polymerization rate in the initial, rapid stages of the reaction. In the mechanism of chain growth, a mono
mer approaches an attached active center and reacts with it. Clearly, an increase in macromolecule length
eventually leads to gradual immurement of active centers, thereby restricting access of the monomer to active
centers. Eventually, chain growth on the ceramic surface becomes diffusionlimited and the polymerization
rate drops. It is of interest to note that the degree of conversion in the initial portion of the kinetic curves is a
linear function of the amount of added ceramic (Fig. 36b). The data presented in Fig. 36b were used to eval
uate the amount of the ceramic that would ensure complete monomer to polymer conversion in the region
of high polymerization rates. As would be expected (Fig. 36a, curve 8), the addition of 90 wt %
Y1Ba2Cu3O7 – x ensures completion of the polymerization at active centers attached to the ceramic surface
[94–98, 109, 112].
The second portion of the kinetic curves 2–7 in Fig. 36a, where the polymerization rate drops sharply
and then gradually increases, represents polymerization in the bulk of the monomer, rather than on the
ceramic surface: reaching a certain size, a polymer shell coiled around a growing active center partially
desorbs from the ceramic surface, exposing the active end of its chain. As such exposed ends accumulate
in the bulk of the monomer, the homogeneous polymerization accelerates, to the extent that there remains

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


34 DAVTYAN et al.

2 Table 9. Effect of filler content on the melting point and enthalpy of fusion of the binder in BPE + Y1Ba2Cu3O6.97
nanocomposites
BPE : Y1Ba2Cu3O6.97
Tm (onset) ΔHm, J/(g BPE Percent crystallinity
weight ratio
90 : 10 107 84 29
97 : 3 107 97 33
99 : 1 105 133 45

no monomer. Note also that the bimolecular chain termination reactions in the bulk have a low rate
because of the low mobility of “exposed” and coiled macromolecules. Clearly, with this mechanism of ini
tiation and chain growth, an increase in ceramic content should lead to an increase in the degree of con
version in the initial stage, where the polymerization process has a high rate, and a drop in the degree of
conversion in the second stage, as observed (Fig. 36a, curves 2–8).
The process follows a different mechanism and has a different topochemistry when radical polymer
ization centers form on the surface of SC ceramic grains and become attached to it.
Polymerization initiation (without AIBN) begins on the ceramic surface and at very high filler contents
in the reaction mixture (90 wt %). The polymerization process has a high rate and seems to be essentially
localized on the ceramic–monomer interface.
Attachment of the ends of macromolecules to the surface of a filler is known to sharply reduce their
mobility and, accordingly, to change the kinetic parameters of the polymerization, in particular, to con
siderably reduce the rate constant of bimolecular chain termination, which is the main cause of the high
rate of the process.
With decreasing ceramic content in the reaction mixture, a second portion appears in the kinetic
curves, pointing to a decrease in the rate of the process. The likely reason for this is that, as the polymer
attached to the ceramic surface accumulates, it restricts access of monomer molecules to polymerization
centers, which thus become occluded (immured) by the forming macromolecules, and the rate of chain
growth also becomes diffusioncontrolled.
Moreover, at certain MMA contents (above 20–25 wt %) the kinetics of the process seem to be more
and more influenced by the transition of radicals from the filler surface to the bulk of the monomer
(through chain transfer to the monomer) and by the homogeneous polymerization initiated by them (sim
ilar to conventional block polymerization). It is reasonable to assume that the rate of block polymerization
should be lower than that of the initial stage of graft polymerization because the diffusion hindrances cre
ated by the filler surface are smaller.

J/(g K)
1.6 1
1.5
1.4
1.3
1.2
1.1
1.0 3
2
0.9
0.8 4
0.7 5
0.6
40 50 60 70 80 90 100 110 120
Temperature, °C

Fig. 40. Temperature dependences of the heat of fusion for SC polymer–ceramic nanocomposites prepared at various initial
forming temperatures (T0) and ceramic : BPE ratios: T0 = (1, 4, 5) 130, (2) 140, and (3) 160°C; Y1Ba2Cu3O6.97 : BPE = (1–3)
90 : 10, (4) 97 : 3, and (5) 99 : 1.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 35

0.5 μm 0.5 μm
(a) (b)

Fig. 41. Micrographs of SC nanocomposites containing (a) PM and (b) PS as binders. Binder : ceramic ratio = 15 : 85.

Although the described qualitative models for the mechanism and topochemistry of MMA on the sur
face of SC ceramics adequately describe observed general trends, this issue remains unresolved, as pointed
out by Davtyan et al. [107, 109, 112].

3.5. Fabrication of Superconducting Polymer–Ceramic Nanocomposites under the Conditions


of the Frontal Polymerization of Mn, Co, Zn, and NiContaining Metal Complex Monomers
in the Presence of Y1Ba2Cu3O6.97
One approach to the search for new SC ceramics with increased SC transition temperatures is doping
(substitution) on particular sites in the crystal lattice of the ceramic. It is therefore reasonable to expect
that one possible way of controlling both the SC transition temperature and the temperature range of the
transition is by using metal complex polymers as binders. To this end, Mn, Co, Zn, and Nicontaining
metallomonomers were polymerized in frontal mode in the presence of varied amounts of Y1Ba2Cu3 O6.97.
At small Y1Ba2Cu3O6.97 additions and temperatures above 100°С, frontal polymerization propagating
downward was only observed at ceramic : metallomonomer weight ratios above 80 : 20. At the same time,
frontal propagation of heat waves was observed at room temperatures, where a wave was initiated from
below and the front propagated upward at various ceramic : metallomonomer ratios.
This was attributed to the gas evolution in the course of frontal polymerization and the inhibiting effect
of some components of the released gas on the polymerization process [108]. It is under upward heat wave
propagation conditions that a number of polymer–ceramic nanocomposites were prepared.

Tg sample−Tg pure, K

10

0 10 30 50 70 90
wt % Y1Ba2Cu3O7 – x relative to the monomer

Fig. 42. Glass transition temperature as a function of ceramic content.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


36 DAVTYAN et al.

0.5 μm 0.5 μm
(a) (b)

0.5 μm 0.5 μm
(c) (d)

Fig. 43. Micrographs of polymer–ceramic nanocomposites containing a UHMWPE binder. Binder : ceramic weight ratio
= (a) 10 : 90, (b) 20 : 80, (c) 59 : 50, and (d) 75 : 25.

The SC and physicomechanical properties of the Mn, Co, Zn, and Nicontaining polymer–
ceramic nanocomposites are summarized in Table 4. It is seen from Table 4 that the SC transition is shifted
to higher temperatures compared to the parent ceramic (Tc = 92 K, Tf = 78 K): Tc by 1–3°С and Tf
by more than 5°С.
It is known from the literature [214] that, above their SC transition, highTc Y1Ba2Cu3O7 – x ceramics
undergo an antiferromagnetic transition, that is, a transition to a spin glass state. Moreover, the antiferro
magnetic and highTc phases are assumed to coexist.
Since Mn, Co, Zn, and Ni are antiferromagnetic metals, it might be expected that the intercalation of
metallopolymer fragments into the interlayer spaces of ceramic grains will further raise Tc compared to
binders free of transition metals. The results show however that the presence of these metals in a polymer
binder has little or no effect on the SC properties of polymer–ceramic nanocomposites.
As shown in the preceding sections, an increase in Tc by 1–3 K can be observed in nanocomposites
containing various binders free of transition metals.

J/(g K)
1
9
3 2
8
7
7
6
5
4
40 60 80 100 120 140 160 180
Temperature, °C

Fig. 44. Temperature dependences of the heat of fusion for SC polymer–ceramic nanocomposites prepared at various ini
tial Y1Ba2Cu3O6.97 ceramic : PP weight ratios: (1) 85 : 15, (2) 70 : 30, and (3) 50 : 50.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 37

0.8 μm 0.8 μm 0.8 μm


(a) (b) (c)

Fig. 45. Micrographs of polymer–ceramic nanocomposites with various Y1Ba2Cu3O6.97 ceramic : PP weight ratios:
(a) 85 : 15, (b) 70 : 30, and (c) 50 : 50.

3.6. Physicomechanical Properties of Superconducting Polymer–Ceramic Nanocomposites


It is known [215] that loading of polymers with fine hard particles typically improves some of the engi
neering parameters of the nanocomposites (stiffness, fracture energy, fracture toughness, heat resistance,
and others), which is commonly attributed to the formation of a special interfacial layer between the filler
and polymer binder.
The special properties of highTc ceramics (layered structure, large surface area of the ceramic grains,
catalytic activity, free oxygen on the surface of the ceramic grains, and others) in comparison with con
ventional fineparticle fillers should have to an unusual effect not only on the formation of interfaces and,
hence, on their physicomechanical properties but also on the SC properties of the polymer–ceramic
nanocomposites.
It is of interest to determine the physicomechanical properties of SC polymer–ceramic nanocompos
ites not only at room temperatures but also at low temperatures, especially below their SC transition tem
perature. Because of this, the tensile strength (σ), elastic modulus (Е), and ultimate strain (ε) of SC poly
mer–ceramic nanocomposites were determined at room temperatures [101–112], and those of materials
containing ultrahighmolecularweight polyethylene as a binder were determined around the αtransition
(193 K) of the polymer matrix and at cryogenic temperatures. At room temperature and 76 K, σ, Е, and
ε were determined in tension [108, 110–112]; at 193 K, they were determined in compression. In the latter
case, the ceramictobinder ratio in the nanocomposites was varied. The σ, Е, and ε data are presented in
Table 5.
Comparison of the strength characteristics for various binders demonstrates that the highest tensile
strength and elastic modulus at almost equal filler contents are ensured by binders based on polyvinyl alco
hol and an ethylene–tetrafluoroethylene copolymer.
It is worth noting that, at low service temperatures, the strength and elastic modulus of polymer mate
rials and particlefilled systems increase, whereas their compliance drops sharply. A criterion for evaluat
ing the performance of polymer materials, especially at low temperatures, is their deformation capability.
In view of this, lowtemperature physicomechanical property measurements were made on nanocompos
ites containing UHMWPE as a binder. Compression tests of the UHMWPEbased nanocomposites at 193

(a) 0.8 μm (b) 0.5 μm

Fig. 46. Electron micrographs of(a) UHMWPE and (b) PP.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


38 DAVTYAN et al.

1
2

3
4

Fig. 47. EPR spectra of superconducting Y1Ba2Cu3O6.97 ceramics and polymer–ceramic nanocomposites with different
binders: (1) PS, 15 wt % relative to the ceramic; (2) PMMA, 15 wt %; (3) UHMWPE, 20 wt %, (4) ST–MMA copolymer,
15 wt %.

K showed that the materials containing 90, 85, and 80 wt % ceramic had a compressive strength of 34, 61,
and 60 MPa, respectively.
It is also of interest to compare the tensile strength at room and cryogenic temperatures. Typical tensile
stress–strain diagrams of Y1Ba2Cu3O6.97 /UHMWPE composites are presented in Fig. 37.
It follows from these data that, at room temperature and equal filler contents, the materials exhibit
highly elastic behavior (Fig. 37a), whereas tension at 77 K leads to brittle fracture in the Hooke’s law
region, where the tensile strain of the SC polymer–ceramic nanocomposites is 0.1%.

3.7. Effects of Particle Size Composition and Filling Fraction on the Physicomechanical
and Superconducting Properties of Polymer–Ceramic Nanocomposites
The key to gaining insight into the processes that determine the properties of particlefilled polymer
materials lies in examining their strength characteristics in relation to the average particle size and size dis
tribution of the filler and binder.
There is currently little data on the influence of the average particle size and particle size distribution
on the properties of polymer composites. Such data are available only in Refs. [216, 217], which assess the
effects of the particle size of the filler and the nature of the binder on the physicomechanical properties of
the composite.
Consider previous data on the effect of filler particle size and concentration on the physicomechanical
properties of polymer–ceramic nanocomposites of Y1Ba2Cu3O6.97 ceramics and divinyl rubber [101, 106,
107, 110]. Table 6 illustrates the effects of particle size and filler content on the tensile strength (σ), elastic
modulus (Е), and ultimate strain (ε) of the nanocomposites.
It is seen from Table 6 that, independent of the grain size of the ceramic, increasing the filler content
leads to an increase in tensile strength and elastic modulus and a drop in ultimate strain. The marked

2, 3
5
4

Fig. 48. EPR spectra of superconducting Y1Ba2Cu3O6.97 ceramics (1) and nanocomposites with a UHMWPE binder: (1)
1, (2) 3, (3) 5, (4) 7, (5) 10, and (6) 20 wt % relative to the ceramic.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 39

Table 10. Superconducting properties of Y1Ba2Cu3O6.97 ceramics and polymer–ceramic nanocomposites


Polymer binder in the nanocomposite
Measure (Y1Ba2Cu3O7 – x : polymer = 85 : 15 by weight)
Supercon
ment
ducting Y1Ba2Cu3O6.97 Poly(StcoMMA)
time, Poly(StcoMMA)
parameter UHMWPE PS + NG2246 (80 : 20 mol %) +
months (60 : 40 mol %)
NG2246
η 0.0189 0.018 0.0197 0.0194 0.0202
0 Tc 92.0 92.8 93.0 92.6 93.4
ΔTc 6.5 6.5 7.0 6.5 9.0
η 0.0185 0.020 0.0185 0.0185 0.0185
6 Tc 91.7 93.8 91.7 92.1 92.2
ΔTc 6.0 6.0 7 7 7
η 0.0185 0.019 0.0181 0.0180 0.0180
12 Tc 91.8 94.8 91.7 91.0 92.0
ΔTc 8 8 9 8 9

increase in σ and Е with increasing filler content suggests that there is good polymer–filler adhesion, with
no binder–filler debonding in the course of mechanical tests.
Further evidence that there is no debonding is provided by the stress–strain diagrams, which show no
yield point.
These results lend support to the conclusion drawn previously that there is rather strong interaction of
the binder with the surface of the ceramic grains and are consistent with the above data on the influence
of ceramic grain size on σ and Е at a constant filler content.
Indeed, it is seen from the data in Table 6 that increasing the average particle size of the filler reduces
both σ and Е. This can be accounted for by the decrease in total binder–filler contact area, which leads
to a reduction in total binder–filler interaction energy and, hence, in ultimate strength and elastic mod
ulus.
It is of interest to note that, with increasing average ceramic grain size, the deformation capability of
the nanocomposites tends to increase.
An increase in particle size probably has an advantageous effect on the cracking resistance of the mate
rial at high strains. There are reports [218] that, like in the above mechanism, rigid filler particles are capa
ble of increasing the ultimate strain of polymer composites relative to the parent polymer.
Also presented in Table 6 are data illustrating the influence of average ceramic grain size on the SC
transition temperature (Tc) and end point (Tf).
Before size separation, the Y1Ba2Cu3O6.97 ceramic had Tc = 93 K and Tf = 87 K. Comparison with the
parent ceramic demonstrates that the DRbased nanocomposites with different ceramic grain sizes differ
in SC properties. The Tc of the nanocomposites with an average ceramic particle size of 5 and 10 μm is 5–
10 K lower than that of the parent ceramic. The transition width (Tc – Tf) also increases. Only at an aver
age particle size of 20–25 μm or above do the nanocomposites have SC properties comparable to or even
better than those of the parent ceramic. This was tentatively attributed to the intercalation of elements of
binder macromolecules into the layered structure of the ceramic grains [98–106].
Thus, the above results lead us to conclude that the key features of the formation of the interface
between the superconducting oxide ceramic and binder, the structure of the binder, and binder–ceramic
adhesion play an important role in determining the superconducting and physicochemical properties of
the polymer–ceramic nanocomposites.
Note that the observed general trends in the formation of the interface in intercalated polymer–
ceramic nanocomposites are similar to those in the formation of the interface (rigid amorphous fraction,
RAF) between a nanoparticle and a polymer macromolecule. This issue was addressed in detail in Refs.
[70, 97, 98, 104, 106, 108–122] and will be analyzed below.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


40 DAVTYAN et al.

Absorbed ethylene, L
16
3
12
2
8 1

0 20 40 60 80
Time, min

Fig. 49. Kinetics of ethylene polymerization with Y1Ba2Cu3O6.97 as a catalyst and Al(C2H5)3 as a cocatalyst.
Al(C2H5)3 : Y1Ba2Cu3O6.97 = (1) 20, (2) 25, and (3) 30.

3.8. Interfacial Phenomena in HighTc Polymer–Ceramic Nanocomposites


Analysis of the literature in the preceding sections has shown an increase in SC transition temperature
by 2–3 K (exemplified by SC polymer–ceramic nanocomposites based on Y1Ba2Cu3O7 – x oxide ceramics
with various polymer binders), which was interpreted in terms of specific interactions of elements of
binder macromolecules with the surface of the ceramic grains. It is reasonable to expect that such inter
action will produce some changes in the arrangement and structure of polymer chains and their confor
mation at the ceramic–polymer interface. So, consider, as an example, the dynamic mechanical and ther
mophysical properties of nanocomposites of ultrahighmolecularweight polyethylene and Y1Ba2Cu3O6.97
ceramic and the results of studies of interfacial phenomena in polymer–ceramic nanocomposites [70, 97,
98, 104, 106, 108–112].

3.9. Dynamic Mechanical Properties of Polymer–Ceramic Nanocomposites


Analysis of previous data on the temperature dependences of the dynamic elastic modulus (Е') and
mechanical loss tangent ( tan δ ) [97, 104, 106, 108, 109, 112] demonstrates that increasing the ceramic con
tent increases the dynamic elastic modulus and its temperature coefficient. As an example, Fig. 38 shows the
temperature dependences of the dynamic elastic modulus for unfilled and particlefilled (15 wt % ceramic)
UHMWPE. Such curves are typical of any rigid fineparticle fillers.
The mechanical loss tangent exhibits a different temperature behavior (Fig. 39, Table 7). The data in
Fig. 39 and Table 7 lead us to conclude that there are two temperature ranges of the largest mechanical
loss: lowtemperature (around –125°С) and hightemperature (~135°С). The lowtemperature transi
tions define the glass transition temperature, and the hightemperature transitions are the melting of the
polymer binder.
Analysis of the data in Fig. 39 and Table 7 indicates that increasing the ceramic content of the nano
composites broadens the mechanical loss peaks, shifts them to higher temperatures, and increases their
height.
It is worth pointing out that such concurrent variations in the mechanical loss characteristics in question is
uncommon for the rigid fineparticle fillers that are typically used in practical applications [219, 220].
Broadening of mechanical loss peaks is commonly attributed to the flaky structure of the filler [221].
The opposite situation occurs most likely in the case under consideration: individual elements of binder
macromolecules are intercalated into the layered structure of the filler and trapped their, producing effects
similar to those of flaky structures. the observed increase in mechanical loss cannot be accounted for by
filler particle agglomeration because filler aggregates form only at high filler contents [222].
The observed increase in mechanical loss is most likely caused by binder adsorption on the filler surface
and the intercalation of macromolecule fragments into the interlayer spaces of the ceramic grains.
Such interaction may lead to changes in the structure of the polymer matrix near the interface and, as
a consequence, to an increase in mechanical loss. It is known [223] that, in a number of cases, a filler shifts
the mechanical loss peak to higher temperatures and increases the glass transition temperature (Tg). It is

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 41

0 50 100
Temperature, K
Fig. 50. SC transition of a polymer–ceramic nanocomposite prepared by gasphase ethylene polymerization in the pres
ence of ceramic Y1Ba2Cu3O6.97. Ceramic : binder ratio = 85 : 15.

believed that the shift is proportional to the surface area of the filler, which can be interpreted in terms of
polymer–filler adsorption interaction.
The nonadditive contribution of the ceramic to the increase in Tg (Table 7) in the case under consid
eration attests not only to adsorption interaction but also, as pointed out above, to intercalation of indi
vidual fragments of UHMWPE macromolecules into the interlayer spaces of the filler grains. Clearly,
these interactions limit the mobility of the macromolecules, changing the packing density of the polymer
chains, their conformation, and, eventually, their morphology near the interface.

3.10. Thermophysical Properties and Morphological Features


of Superconducting Polymer–Ceramic Nanocomposites
To further validate the conclusions drawn in the preceding section, consider results of differential scan
2 ning calorimetry studies [108, 112] in which the melting point (Tm) and the enthalpy of fusion (ΔHm) of
nanocomposites of a UHMWPE binder and Y1Ba2Cu3O6.97 ceramic were directly measured in a wide
composition range. The results obtained by Davtyan et al. [108, 112] are summarized in Table 8. It can be
2 seen that the enthalpy of fusion increases with increasing filler content.

2 The observed increase in enthalpy is related to either the degree of crystallinity (Tables 8, 9) or changes
in the morphology of the binder near the interface. One however cannot ascertain from the data in ques
tion which of these factors plays a key role. To this end, the effect of Y1Ba2Cu3O6.97 ceramic content on
the heat of fusion was studied by DSC techniques [97, 104, 108, 112] under temperature scan conditions,
and polymer–ceramic nanocomposites were examined in detail by electron microscopy.
Figure 40 shows the temperature dependences of the heat of fusion for SC polymer–ceramic nano
composites prepared at various initial forming temperatures and containing various amounts of BPE.
It is seen from the curves in Fig. 40 that the highest values of the heat of fusion (Fig. 40, curves 1–4) of
the SC polymer–ceramic materials are essentially independent of the initial forming temperature (Fig. 40,
curves 1–3) but depend rather strongly on the ceramic content of the composites (Fig. 40, curves 4, 5). The
melting points and enthalpies of fusion evaluated from the data in Fig. 40 are listed in Table 9.

Table 11. Effect of nanocomposite composition on the superconducting transition temperature and transition width
(Tc–Tf)
Ceramic : polyethylene
Tcl, К Tcl–Tfl, К Tcm, К Tcm–Tfm, К Tch, К Tch–Tfh, К
weight ratio
100 : 0 10 4.5 ∼30 ∼10 46 ∼30
80 : 20 10 4.5 ∼50 10 70 ∼50
50 : 50 79 ∼63

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


42 DAVTYAN et al.

Tc
82

78

74

70
0 4 8 12
Time, months

Fig. 51. Variation in Tc during aging of an SC nanocomposite containing 20 wt % polyethylene.

2 Thus, both the enthalpy and degree of crystallinity increase with filler content. Therefore, by analogy
with the above case, it is reasonable to assume that the rather large increase in ΔHm (relative to the BPE)
may be due to two causes:
1. An increase in ceramic content leads to an increase in the degree of crystallinity.
2. Intercalation of fragments or individual elements of binder macromolecules into the interlayer
spaces of the ceramic grains leads to changes in the morphology of the branched polyethylene on the
ceramic–binder interface, which appears more likely [108, 109, 111, 112].
Indeed structural studies of SC polymer–ceramic nanocomposites by scanning electron microscopy
demonstrate that, with both amorphous and crystalline polymers, the ceramic grains become completely
and uniformly covered with polymer binders (Fig. 41), which can be interpreted as evidence that there is
rather strong ceramic–polymer interfacial interaction.
The interaction of PM and PS binder macromolecules with the surface of Y1Ba2Cu3O6.97 ceramic
grains may contribute to the formation of a rigid amorphous fraction (RAF) in the polymer binder.
By analogy with previous studies [13, 67–70, 138, 146, 148], the presence of a RAF on the surface of
Y1Ba2Cu3O6.97 ceramic grains may lead to an increase in the glass transition temperature (Tg) of the binder.
Indeed, as seen in Fig. 42, the glass transition temperature of an SC composite containing PMMA as a
binder increases by more than 15°С with increasing ceramic content.

Table 12. Aging effect on the superconducting properties of Mn, Co, Zn, and Nicontaining SC nanocomposites
Composition of the nanocomposite
Pressing
Y1Ba2Cu3O6.97 metal complex monomer Metal Tс/Tс1, K Tf/Tf1, K
time, min
g wt % g wt %
0.293 43 0.388 57 Mn 10 95/96 87/89
0.396 50 0.396 50 Mn 5 94/96 85/87
0.90 70 0.290 30 Mn 5 94/95 84/88
0.518 73 0.196 27 Mn 5 93/96 85/90
0.900 70 0.390 30 Mn 5 95/96 85/91
0.416 67 0.209 33 Mn 10 94/96 83/86
0.552 78 0.156 22 Mn 5 95/96 85/88
0.325 51 0.318 49 Co 5 93/95 83/85
0.432 60 0.283 40 Co 5 92/96 84/89
0.503 70 0.228 30 Co 2 92/95 84/87
0.486 70 0.208 30 Ni 5 95/96 83/85

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 43

Nanocomposites with UHMWPE and PP binders contain fibrous structures, independent of ceramic
content. Figure 43 shows electron micrographs of materials containing PE as a binder. There are fibrous
structures atypical of polyethylene.
As mentioned above, fibrous structures may result from the intercalation of fragments of UHMWPE
macromolecules into the layered structure of ceramic grains. Such “attachment” may influence the
mobility of the polyethylene macrochains, reducing their flexibility. Therefore, the crystallization of mac
romolecules linked in this way gives rise to intermolecular interaction between them.
Of special note is that, in the case of polymer–ceramic nanocomposites containing PP as a binder, the
effects described above are more pronounced. Indeed, in the temperature dependence of the heat of fusion
in Fig. 44, the peaks are split into two components.
The splitting is most likely caused by the presence of two structures in the polymer–ceramic nanocompos
ites containing PP. This assumption can be validated by electron microscopy data. Indeed, the micrographs of
PPcontaining materials (Fig. 45) show considerably more fibrous structures than in the case of PE.
It follows from Fig. 45 that at an Y1Ba2 Cu3 O6.97 : PP ratio of 85 : 15 or above, there are double and
triple fibrous structures. The large number of fibrous structures is probably responsible for the splitting of
the peaks at the melting point (Fig. 44) in S polymer–ceramic nanocomposites with an isotactic propy
lene binder.
The ends of such crystalline fibrous structures attach to ceramic particles, thereby suggesting that they
can be incorporated into the interlayer spaces of the ceramic grains (Figs. 43a, 45a). It should be noted
that no such fibrous structures were detected in unfilled crystalline polymer binders (Figs. 46a, 46b).

3.11. Effect of Filling Fraction on the Valence State of Copper in Polymer–Ceramic Nanocomposites
HighTc oxide ceramics are known to possess intrinsic localized magnetic moments, which produce a
Cu2+ EPR signal. It is worth noting however that Y1Ba2Cu3O7 – x ceramics contain two types of copper
atoms: Cu2+(1) and Cu2+(2). The former atoms reside in CuO chains along the b axis, and the latter atoms
are located in CuО2 planes in the ab plane. The origin of the EPR signal from Y1Ba2Cu3O7 – x ceramics was
unclear for a long time [224]. Cu2+ EPR signal measurements, in parallel with Cu Cu2+edge Xray
absorption near edge structure (XANES) measurements for the same samples [227] and analysis of the
influence of Fe substitution for Cu2+ (1) in Y1Ba2Cu3O7 – x on the Cu2+ EPR signal intensity, showed [228]
that the EPR signal was due to the Cu2+(1) atoms in the chains, and not to the Cu2+(2) atoms in the planes.
EPR data for polymer–ceramic nanocomposites demonstrate that the Cu2+(1) EPR signal depends on the
nature of the binder.
Figure 47 shows the Cu2+(1) EPR spectra of Y1Ba2Cu3O6.97 ceramics and polymer–ceramic nanocom
posites with different polymer binders: polystyrene, polymethyl methacrylate, and polyethylene.
It follows from curves 1–4 in Fig. 47 that the addition of the polymers changes the valence state of
Cu2+(1), which is direct evidence of intermolecular interaction between the Y1Ba2Cu3O6.97 ceramic grains
and elements of polymer chains. This interaction can be explained by the intercalation of individual ele
ments or fragments of binder macromolecules into the layered structure of the ceramic. As a result of the
intercalation, the orbital of the unpaired Cu2+(1) electron overlaps with the orbitals of the corresponding
elements in the binder macrochains, leading to changes in the valence state of Cu2+(1) and, eventually, in
the Cu2+(1) EPR signal intensity.
It is of interest to find out whether or not the Cu2+(1) EPR signal intensity depends on filler content,
that is, on the ceramic : binder ratio.
To this end, polymer–ceramic nanocomposites were produced with Y1Ba2Cu3O6.97 : UHMWPE
weight ratios of 100 : 0, 99 : 1, 97 : 3, 93 : 7, 90 : 10, and 80 : 20. The EPR results for these materials are
presented in Fig. 48. Curves 1–6 in Fig. 48 demonstrate that the EPR signal intensity depends on binder
content. It is of interest to note that the largest variation from the pure ceramic (Fig. 48, curve 1) occurs
at low binder contents (Fig. 48, curves 2, 3). Increasing the percentage of the binder reduces the signal
intensity (Fig. 48, curves 3–6).
Thus, the electron microscopy and Cu2+(1) EPR data for the polymer–ceramic nanocomposites pro
vide direct evidence of intermolecular interaction between the polymer binder and the surface of ceramic
grains through the intercalation of elements of binder macromolecules into the interlayer spaces of the
ceramic grains.
Taking into account the above data on the dynamic mechanical properties of superconducting UHM
WPEbased nanocomposites in a wide temperature range, we conclude that the key features of the forma
tion of the ceramic–binder interface play an important role in determining the superconducting and
physicomechanical properties of the nanocomposites.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


44 DAVTYAN et al.

The electron microscopy and Cu2+(1) EPR data provide direct evidence for the intercalation of frag
ments of macromolecules into the interlayer spaces of the ceramic, resulting in the formation of nano
structures in the ceramic grains.

3.12. Aging of Y1Ba2Cu3O6.97 Ceramics and Nanocomposites Containing Different Binders


The properties of highTc Y–Ba–Cu–O superconductors are known to be determined by both the net
oxygen content and the degree of order in the oxygen distribution over crystallographic sites [229–237].
The present generation of SC ceramics contain an unstable phase. The lattice instability, accompanied by
the generation of ordered vacancies in the Cu–O chains, is caused by oxygen deficiency in the unit cell of
the material. Therefore, it cannot be ruled out that the lattice instability of SC Y1Ba2Cu3O7 – х ceramics
increases over time, which may lead to changes in their superconducting properties (Tc and ΔTc) and
5 orthorhombic lattice distortion.
To address this issue, polymer–ceramic nanocomposites were produced by various techniques [94, 98,
100, 101, 107, 109, 111, 112], and the superconducting properties of Y1Ba2Cu3O6.97 ceramics and the
polymer–ceramic nanocomposites were monitored during storage for a year. The SC ceramics and poly
mer–ceramic composites were stored in air at room temperature. The SC ceramics and composites were
characterized by Xray diffraction on a DRON2.0 diffractometer (λCuKα) in the angular range 2θ = 15°–
130° at room temperature. The grains in the Y1Ba2Cu3O6.97 ceramic were found to have (110) preferential
alignment, as was evidenced by the fact that the (110) line had the highest intensity, which remained con
stant over time (Table 10). The sample was oriented along (006), and the intensity of this reflection did not
vary over time. In addition, it contained 2% Y1BaCuO5 (211), a semiconducting phase.
The Y1Ba2Cu3O6.97 samples were found to have rather stable superconducting properties. Slight
changes in their SC properties were irreversible and were probably related to autooscillating processes
typical of structurally unstable electron systems in solids, rather than to structural aging. It seems likely
that the semiconducting phase Y1BaCuO5 in Y1Ba2Cu3O6.97 is an oxygen transport phase [238] and, when
present in the ceramic in certain concentrations, acts to stabilize its SC properties by reducing the con
centration of vacant oxygen sites.
As shown in the preceding sections, the slight increase in Tc immediately after the pressing of the poly
mer–ceramic composites (Table 10) is due to the intercalation of fragments or individual elements of
binder macromolecules into the interlayer spaces of the ceramic grains during hot pressing. At the same
time, analysis of ample data [94, 98, 100, 101, 107, 109, 111, 112] on the aging kinetics of SC polymer–
ceramic nanocomposites indicates that the evolution of their SC properties depends as well on the chem
ical composition of the polymer binder. The SC properties of nanocomposites containing ST–MMA
copolymers as binders degrade, whereas those of UHMWPEmatrix composites improve. The observed
increase in the SC transition temperature of polyethylenematrix nanocomposites is most likely related to
processes that take place after the fabrication of the polymer–ceramic nanocomposites and is determined
by the interaction of elements of the polymer binder with the surface of the ceramic grains, to the point of
intercalation of such elements into the interlayer spaces of the ceramic grains.
Based on the data in Table 10, it is reasonable to assume that the interaction of polymer binder macromol
ecules with the surface of the ceramic grains by the intercalation mechanism persists at room temperature but
becomes weaker. Because of this, the UHMWPEmatrix composites have a Tc increased by 1–1.5 K.
As mentioned above, aging reduces the Tc of the SC composites containing ST–MMA copolymers as
binders by 1–1.5 K and increases their ΔTc . At the same time, the presence of NG2246 antioxidant addi
tions has no effect on these properties. This can be accounted for by the strong tendency for ST–MMA
copolymers to decompose when exposed to UV radiation.
More detailed data on the increase in Tc during aging of polymer–ceramic nanocomposites were
obtained for polyethylene binders. To this end, similar studies were carried out using samples prepared by
gasphase ethylene polymerization in the presence of Y1Ba2Cu3O6.97 ceramics.

3.13. Superconducting Polymer–Ceramic Nanocomposites Produced


by GasPhase Ethylene Polymerization and Their Aging
Perovskite oxide ceramics have long been known to possess catalytic properties [209–211]. A natural
question in this context is whether the surface of such ceramics can be catalitically activated in order to
use them as catalysts for the polymerization of gaseous monomers (ethylene, propylene, and others).
5 It seems likely that the layered structure of the crystalline orthorhombic phase allows ethylene (or other
olefins) to be coordinated on the surface of Y1Ba2Cu3O7 – x ceramic grains. Therefore, when the reaction

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 45

medium contains cocatalysts (alkylaluminum), polymerization is possible. To this end, [111, 112],
Y1Ba2Cu3O6.97 powder less than 50 μm in particle size was heattreated at 497°С for 4 h and then cooled
to room temperature in a dry nitrogen atmosphere. Part of the ceramic thus heattreated was used to
determine its SC properties, and the rest was tested as a catalyst for polymerization. To a polymerization
reactor (filled with hexane) was added an appropriate amount of the ceramic. Next, ethylene was intro
duced and an alkylaluminum as a cocatalyst was injected with vigorous stirring (100 rpm). Ethylene
absorption continued for 3 h (Fig. 49), providing clear evidence of ethylene polymerization.
In this way, using powdered Y1Ba2Cu3O6.97 ceramics and gasphase ethylene polymerization in heptane,
we obtained a number of composites differing in polyethylene content.
αС susceptibility measurements [100, 111, 112] showed that the SC transition curves of all the nano
composites (Fig. 50) had three features corresponding to three SC transitions: a lowtemperature transi
tion (Tlc ~ 5 K), independent of the composition of the composite; a mediumtemperature transition
(Tmc); and a relatively hightemperature one (Thc). The last two transitions depended significantly on the
polyethylene content of the composites. The three transition temperatures, Tlc, Tmc, and Thc, are indicated
in Table 11.
The observed stepwise SC transition can be tentatively accounted for by variations in the degree of deg
radation from region to region and, accordingly, to a reduction in intercalation rate, which depends on the
degree of amorphization of the ceramic grains.
It seems likely that ethylene polymerization occurs in part directly in the interlayer spaces of the
ceramic grains. The rate of this process depends on the degree of amorphization of the different fractions
of the Y1Ba2Cu3O6.97 ceramic.
It is of interest to note that the superconducting transition onset (Tch) and transition width (Tch – Tcm)
depend significantly on the composition of the nanocomposite. As mentioned above, the observed
increase in Tch and Tch – Tcm is attributable to the high rate of the polymerization in the interlayer spaces
of the ceramic grains.
The aging kinetics of an SC nanocomposite containing 20% PE were examined every two months (Fig. 51)
[109, 111, 112].
These data demonstrate that Tc varies at an increasing rate. It seems likely that initial incorporation of
fragments of macromolecules into the layered structure of the ceramic grains subsequently facilitates this
process. In addition, it follows from Fig. 51 that no saturation was reached even after eight months of
aging: the intercalation process continued at a slow rate.
These results provide convincing evidence in favor of the assumption that, even at room temperatures,
fragments of macromolecules are intercalated into the interlayer spaces of the ceramic grains.

3.14. Aging of Superconducting Polymer–Ceramic Nanocomposites Produced by the Frontal Polymerization


of Metal Complex Monomers in the Presence of Ceramic Y1Ba2Cu3O6.97
Mn, Co, Zn, and Nicontaining SC polymer–ceramic composites containing 30% metallopoly
mers were tested for aging. The SC properties of these materials are summarized in Table 12. These data
demonstrate that aging of the Mn, Co, Zn, and Nicontaining polymer–ceramic nanocomposites also
increases their superconducting transition temperature (Tc1) by 1–3 K and reduces their transition width
(Tc1 – Tf) by 4–6 K. No explanation for the rather large decrease in ΔTc was proposed by Davtyan et al.
[109, 111, 112].
Thus, analysis of the results reported in Refs. [91–113] leads us to conclude that activation of the sur
face of oxide (Y1Ba2Cu3O6.97) ceramic grains allows such ceramics to be used as catalysts for gasphase
polymerization of olefins. Studies of the aging kinetics of SC –ceramic nanocomposites [109, 111, 112]
demonstrate that, even at room temperatures, fragments of binder macromolecules are intercalated into
the interlayer spaces of the ceramic, leading to the formation of unusual nanostructures in the ceramic
grains.

REFERENCES
1. Barber, A.H., Cohen, S.R., and Wagner, H.D., Appl. Phys. Lett., 2003, vol. 82, p. 4140.
2. Barber, A.H., Cohen, S.R., Kenig, S., and Wagner, H.D., Compos. Sci. Technol., 2004, vol. 64, p. 2283.
3. Barber, A.H., Cohen, S.R., and Wagner, H.D., Phys. Rev. Lett., 2004, vol. 92, article 186 103.
4. Kuoa, M.C., Tsaia, C.M., Huanga, J.C., and Chena, M., Mater. Chem. Phys., 2005, vol. 90, p. 185. 5.
Zhu, Y.Q., Sekine, T., Li, Y.H., Fay, M.W., Zhao, Y.M., Patrick, P.C.H., et al., J. Am. Chem. Soc., 2005,
vol. 127, p. 16263.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


46 DAVTYAN et al.

6. Buchachenko, A.L., Usp. Khim., 2006, vol. 75, pp. 3–26.


7. Wu, M. and Shaw, L., J. Appl. Polym. Sci., 2006, vol. 99, p. 477. 8. Munaro, M., Brandt, T.R., and Leni, C.A.,
Morphological Characterization of LDPE / HDPE Blends by Dynamical Mechanical Analysis, Proc. World
Polymer Congr.–Macro 2006, 41st Int. Symp. on Macromolecules, Rio de Janeiro, 2006.
9. Vera Lúcia da Cunha Lapa, Patrícia Davies de Oliveira, John Duncan, Leila Léa Yuan Visconte, and Regina
Celia Reis Nunes, Dynamic Properties in NBR/Cel II Nanocomposites, Proc. World Polymer Congr.–Macro
2006, 41st Int. Symp. on Macromolecules, Rio de Janeiro, 2006.
10. Kontou, E. and Niaounakis, M., Polymer, 2006, no. 47, p. 1267.
11. Zeng, X.F., Wang, W.Y., and Wang, G.Q.J.F, J. Mater. Sci., 2008, vol. 43, p. 3505.
12. Tonoyan, A.O., Ketyan, A.G., Shik, K., and Davtyan, S.P., Khim. Zh. Arm., 2009, vol. 62, nos. 1–2, p. 201.
13. Davtyan, S.P., Berlin, A.A., Shik, K., Tonoyan, A.O., and Ragovina, S.Z., Ross. Nanotekhnol., 2009, vol. 4,
nos. 7–8, p. 489.
14. Tonoyan, A.O., Ketyan, A.G., Zakaryan, A.O., Sukiasyan, A.A., Sukiasyan, Zh.K., and Davtyan, S.P., Khim.
Zh. Arm., 2010, vol. 63, p. 193.
15. Yonghong, R. and Sie Chin Tjong, ePolymers, 2010, no. 109.
16. Rabova, V. and Hron, P., ePolymers, 2011, no. 091.
17. Yingchun Li, Guosheng Hu, and Bin He, ePolymers, 2011, no. 060.
18. Yu, Y.Y. and Chen, W.C., Mater. Chem. Phys., 2003, vol. 82, p. 388.
19. Liu, T., Burger, C., and Chu, B., Prog. Polym. Sci., 2003, vol. 28, p. 5.
20. Ajayan, P.M., Schadler, L.S., and Braun, P.V., Nanocomposite Science and Technology, Weinheim: WileyVCH,
2003.
21. Cao, G., Nanostructures and Nanomaterials: Synthesis, Properties and Applications, London: Imperial College
Press, 2004.
22. The Chemistry of Nanomaterials, Rao, C.N.R., Muller, A., and Cheetman, A.K., Eds., Weinheim: Wiley–VCH,
2004.
23. Fujimura, K., Ohkoshi, Y., Nagura, M., Akamatsu, K., Deki, S., and Gotoh, Y., Mater. Chem. Phys., 2004,
vol. 83, p. 54.
24. Nanoparticles, Schmid, G., Ed., New York: Wiley–Interscience, 2004.
25. Springer Handbook of Nanotechnology, Bhushan B, Ed., Berlin: Springer, 2004.
26. Pomogailo, A.D. and Kestelman, V., Metallopolymer Nanocomposites, Berlin: Springer, 2005.
27. Liu, P., Eur. Polym. J., 2005, vol. 41, p. 2693.
28. AbdolReza Hajipour and Sakineh Habibi, ePolymers, 2010, no. 012.
29. Yibing Cai, Xiaolin Xu, Qufu Wei, Huizhen Ke, Wang Yao, Hui Qiao, Chuilin Lai, Guangfei He, Yong Zhao,
and Hao Fong, ePolymers, 2011, no. 018.
30. Agabekov, V., Ivanova, N., Dlugunovich, V, and Vostchula, I., Optical Properties of Polyvinyl Alcohol Films,
Modified with Silver Nanoparticles, Special Issue: Synthesis, Properties, and Applications of Polymeric Nano
composites, J. Nanomater., 2012 (in press).
31. Fan, X., Wang, F., Lu, X., Wang, M., Qiu, J., and Kawamoto, Y.T., Mater. Chem. Phys., 2003, vol. 82, p. 38.
32. Wang, Z., Wang, J., and Zhang, H., Mater. Chem. Phys., 2004, vol. 87, p. 44.
33. Peng, T., Huajun, L., Yang, H., and Ya, C., Mater. Chem. Phys., 2004, vol. 85, p. 68.
34. Lu, C.H., Hsu, W.T., Huang, C.H., Godbole, S.V., and Cheng, B.M., Mater. Chem. Phys., 2005, vol. 90,
p. 62.
35. Vollath, D., Szabo, D.V., and Schlabach, S., J. Nanopart. Res., 2004, vol. 6, p. 181.
36. Aharon, E., Albo, A., Kalina, M., and Frey, G.L., Adv. Funct. Mater., 2006, vol. 16, p. 980.
37. Zhang Xiaozhou, Jian Xigao, and Zu Liwu. Materials and Devices Research of PPV–ZnO Nanowires for
Heterojunction Solar Cells, Special Issue: Synthesis, Properties, and Applications of Polymeric Nanocompos
ites, J. Nanomater., 2012 (in press).
38. Bissessur, R., Gallant, D., and Brüening, R., Mater. Chem. Phys., 2003, vol. 82, p. 316.
39. Tsunoyama, H., Sakurai, H., Ichikuni, N., Negishi, Y., and Tsukuda, T., Langmuir, 2004, vol. 20, p. 11 293.
40. Sonawane, R.S., Kale, B.B., and Dongare, M.K., Mater. Chem. Phys., 2004, vol. 85, p. 52.
41. Chowdhury, D., Paul, A., and Chattopadhyay, A., Langmuir, 2005, vol. 21, p. 4123.
42. Einage, H. and Harada, M., Langmuir, 2005, vol. 21, p. 2578.
43. Bertoncello, P., Notargiacomo, A., and Nicolini, C., Langmuir, 2005, vol. 21, p. 172.
44. Ozturk, O., Black, T.J., Perrine, K., Pizzolato, K., Parsons, F.W., Ratliff, J.S., et al., Langmuir, 2005, vol. 21,
p. 998.
45. Singh, A. and Chandler, B.D., Langmuir, 2005, vol. 21, p. 10776.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 47

46. Torrens, F. and Castellano, G., Fractal Dimension of ActiveSite Models of Zeolite Catalysts, J. Nanomater.,
2006, article ID 17 052.
47. Srinivasan, S.S., Wade, J., and Stefanakos, E.K., Synthesis and Characterization of Photocatalytic TiO2–
ZnFe2O4 Nanoparticles, J. Nanomater., 2006, article ID 45 712.
48. Zhong, K., Jin, P., and Chen, Q., Ni Hollow Nanospheres: Preparation and Catalytic Activity, J. Nanomater.,
2006, article ID 37 375. doi 10.1155/JNM/2006/37375
49. Sesha S. Srinivasan, Jeremy Wade, and Elias K. Stefanakos, Visible Light Photocatalysis via CdS/TiO2 Nano
composite Materials, J. Nanomater., 2006, article ID 87 326. doi 10.1155/JNM/2006/87326
50. Foll, H., Carstensen, J., and Frey, S., Porous and Nanoporous Semiconductors and Emerging Applications,
J. Nanomater., 2006, article ID 91 635. doi 10.1155/JNM/2006/91635
51. Yi Yang, XinJie Yu, YaXue Wang, and Jun Wang, Assessment on the Dispersibility of Nanocatalyst in Ener
getic Polymer by Employing Relative Standard Deviation Method, Special Issue: Synthesis, Properties, and
Applications of Polymeric Nanocomposites, J. Nanomater., 2012 (in press).
52. Liu, X., Lee, C., Zhou, C., and Han, J., Appl. Phys. Lett., 2001, vol. 79, p. 3329.
53. Shim, M., Javey, A., Kam, N.W.S., and Dai, H., J. Am. Chem. Soc., 2001, vol. 123, p. 11 512.
54. Dai, H., Acc. Chem. Res., 2002, vol. 35, p. 1035.
55. Abes, J.I., Cohen, R.E., and Ross, A., Chem. Mater., 2003, vol. 15, p. 1125.
56. Vasques, C.T., Domenech, S.C., Barreto, P.L.M., and Soldi, V., ePolymers, 2010, no. 026.
57. Hirmizia, M.Abu., Bakara, W.L., Tana, N.H.H., Bakara, A., Ismaila, J., and Seeb, C.H., Electrical and Ther
mal Behavior of Copper–Epoxy Nanocomposites Prepared via Aqueous to Organic Phase Transfer Technique,
Special Issue: Synthesis, Properties, and Applications of Polymeric Nanocomposites, J. Nanomater., 2012 (in
press).
58. He, B., Suna, W., Wang, M., Liu, S., and Shen, Z., Mater. Chem. Phys., 2004, vol. 84, p. 140.
59. Jafari Nejad, Sh., ePolymers, 2012, no. 025.
60. Najeh, I., Dahman, H., Mansour, N.B., and Mir, L.E., Electrical Investigations and Dielectric Properties of
Nanoporous Carbon, Special Issue: Synthesis, Properties, and Applications of Polymeric Nanocomposites,
J. Nanomater., 2012 (in press).
61. Olejnik, R., Slobodian, P., Riha, P., and Saha, P., An ElectricallyConductive and Organic Solvent Vapors
Detecting Composite Composed of an Entangled Network of Carbon Nanotubes Embedded in Polystyrene,
Special Issue: Synthesis, Properties, and Applications of Polymeric Nanocomposites, J. Nanomater., 2012 (in
press).
62. Ismail Ab Rahmana and Vejayakumaran Padavettanb, Synthesis of Silica Nanoparticles by Sol–Gel; Size
Dependent Properties, Surface Modification and Applications in Silica–Polymer Nanocomposites. Special
Issue: Synthesis, Properties, and Applications of Polymeric Nanocomposites, J. Nanomater., 2012 (in press).
63. Renhua Zheng, Huajiang Jiang, Haichang Guo, and Weilin Sun, ePolymers, 2011, no. 086.
64. Casalboni, M., Dominici, L., Foglietti, V., Michelotti, F., Orsini, E., Palazzesi, C., Stella, F., and
Prosposito, P., Bragg Grating Optical Filters by UV Nanoimprinting, Special Issue: Synthesis, Properties, and
Applications of Polymeric Nanocomposites, J. Nanomater., 2012 (in press).
65. Savva, I., Krekos, G., Taculescu, A., Marinica, O., Vekas, L., and KrasiaChristoforou, T., Fabrication and
Characterization of MagnetoResponsive Electrospun Nanocomposite Membranes Based on Methacrylic
Random Copolymers and Magnetite Nanoparticles, Special Issue: Synthesis, Properties, and Applications of
Polymeric Nanocomposites, J. Nanomater., 2012 (in press).
66. Xie Minqiang, Xu Yiming, Liu Jie, Zhang Tao, Zhang Hongzheng, Preparation and Characterization of Folate
Targeting Magnetic Nanomedicine Loaded with Cisplatin, Special Issue: Synthesis, Properties, and Applica
tions of Polymeric Nanocomposites, J. Nanomater., 2012 (in press).
67. Sargsyan, A.G., Tonoyan, A.O., Davtyan, S.P., and Schick, C., Eur. Polym. J., 2007, no. 8, p. 3113.
68. Sargsyan, A.G., Tonoyan, A.O., Davtyan, S.P., and Schick, C., 9th Eur. Symp. in Thermal Analysis and Calo
rimetry, Krakow, 2006, p. 37.
69. Sargsyan, A.G., Tonoyan, A.O., Davtyan, S.P., and Schick, C., NATAS Notes, 2007, vol. 39, no. 4, p. 6.
70. Davtyan, S.P. and Tonoyan, A.O., Osnovy nanotekhnologii. Nanochastitsy i polimernye nanokompozity (Princi
ples of Nanotechnology: Nanoparticles and PolymerMatrix Composites), Yerevan: Nauka RA, 2011.
71. Bednorz, J.G. and Muller, K.A., Z. Phys., 1986, vol. 64, no. 2, p. 189.
72. Bednorz, J.G. and Muller, K.A., Rev. Mod. Phys., 1988, vol. 60, p. 585.
73. Sharma, R.F., Reddy, Y.S., and Pzamana, S., Phys., 1988, vol. 30, p. 181.
74. Weaver, J.H., Meyer, H.M.III., Wagener, T.J., Hill, D.M., Gao, Y., Peterson, D., Fisk, Z., and Arko, A.J., Phys.
Rev. B: Condens. Matter Mater. Phys., 1988, vol. 38, p. 4668.
75. Meyer, H.M. and Hill, D.M., Phys. Rev. B: Condens. Matter Mater. Phys., 1988, vol. 38, p. 6500.
76. Meyer, H.M., Hill, D.M., Wagener, T.J., Weaver, J.H., Gallo, C.F., and Goretta, K.C., J. Appl. Phys., 1989,
vol. 65, p. 3130.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


48 DAVTYAN et al.

77. Romana, L.T. and Wilshaw, P.R., Supercond. Sci. Technol., 1989, vol. 6, p. 285.
78. Lindbery, P.A. and Shen, Z., Phys. Rev. B: Condens. Matter Mater. Phys., 1989, vol. 39, p. 2890.
79. Hikata, T. and Sato, K.J., J. Appl. Phys., 1989, vol. 28, no. 1, p. 182.
80. Gagafarov, S.F. and Dzhafarov, T.D., Zh. Tekh. Fiz., 1990, vol. 16, no. 9, p. 59.
81. Asoka, P.S. and Mahumini, K., J. Appl. Phys., 1990, vol. 67, no. 6, p. 3184.
82. Dotsenko, V.I., Braude, V.I., Ivaechenko, L.G., and Kislyak, I.F., Fiz. Nizk. Temp., 1996, vol. 22, no. 10, p. 222.
83. Frase, K.G., Liniger, E.G., and Clarke, D.R., Adv. Ceram. Mater., 1987, vol. 2, no. 3B, p. 698.
84. Goto, T. and Kada, M., J. Appl. Phys., 1987, vol. 26, no. 9, p. 1524.
85. Golyamina, E.M. and Lykov, A.N., Sverkhprovodimost: Fiz., Khim., Tekh., 1989, vol. 2, no. 6, p. 51.
86. Gummins, T.R., Egdell, R.G., and Georgiadis, G.C., J. LessCommon Met., 1990, vols. 164–165, p. 1149.
87. Pulyaeva, I.V., Puzanova, A.A., Eksperindova, L.P., Nekrasova, L.N., et al., Sverkhprovodimost: Fiz., Khim.,
Tekh., 1993, vol. 6, p. 1430.
88. Pulyaeva, I.V., Eksperindova, L.P., Kovtun, E.D., Mateichenko, V.P., et al., Inorg. Mater., 1996, vol. 32, p. 788.
89. Haupt, S.G., Riley, D.R., Jones, C.T., Zhao, J., and McDevitt, J.T., J. Am. Chem. Soc., 1993, vol. 115, p. 1196.
90. Haupt, S.G., Riley, D.R., Zhao, J., and McDevitt, J.T., J. Phys. Chem., 1993, vol. 97, p. 7796.
91. Torosyan, A.A, Nersesyan, N.D., Peresada, A.G., Davtyan, S.P., Borovinskaya, I.P., and Merzhanov, A.G.,
USSR Inventor’s Certificate no. 4 737 132, 1989.
92. Torosyan, A.A., Nersesyan, M.D., Peresada, A.G., Davtyan, S.P., Borovinskaya, I.P., and Merzhanov, A.G.,
USSR Inventor’s Certificate no. 4 749 647, 1989.
93. Gyulumyan, X.N. and Davtyan, S.P., USSR Inventor’s Certificate no. 1 831 197, 1990.
94. Arakelova, E.R., Bagdasaryan, A.E., Mirzoyan, G.N., Tonoyan, A.O., and Davtyan, S.P., Khim. Zh. Arm.,
1997, vol. 50, nos. 1–2, p. 24.
95. Davtyan, S.P., Airapetyan, S.M., and Tonoyan, A.O., Preparation and Properties of HighTc Polymer–
Ceramic Nanocomposites, in Konversionnyi potentsial Armenii i programmy MNTTs (Conversion Potential of
Armenia and ISTC Programs), Yerevan, 2000, p. 253.
96. Tonoyan, A.O. and Davtyan, S.P., HighTc Polymer–Ceramic Nanocomposites: Frontal Polymerization of
Crystalline Monomers with the Aim of Producing Bulk and ParticleFilled Polymer Materials, in Konversion
nyi potentsial Armenii i programmy MNTTs (Conversion Potential of Armenia and ISTC Programs), Yerevan,
2000, p. 162.
97. Davtyan, S.P., Hayrapetyan, S.M., and Tonoyan, A.O., High Temperature SC Polymer–Ceramic Composites
and Their Morphological Peculiarities, in Chemistry in Armenia on the Threshold of the XXI Century, Yerevan,
2000, p. 41.
98. Hayrapetyan, S.M., Tonoyan, A.O., Arakelova, E.A., and Davtyan, S.P., Producing High Temperature SC
Polymer–Ceramic Composites and Their Properties in Chemistry in Armenia on the Threshold of the XXI Cen
tury, Yerevan, 2000, p. 56.
99. Davtyan, S.P., Hayrapetyan, S.M., Tonoyan, A.O., Hasratyan, A.K., Israelyan, V.R., and Hovnanyan, K.O.,
Int. Symp. High Temperature SC Polymer–Ceramic Composites and Their Morphological Peculiarities, Brno,
2000, p. 125.
100. Tonoyan, A.O., Davtyan, S.P., Martirosyan, S.A., and Mamalis, A.G., J. Mater. Process. Technol., 2001,
vol. 108, p. 201.
101. Airapetyan, S.M., Tonoyan, A.O, Arakelova, E.R., and Davtyan, S.P., Vysokomol. Soedin., Ser. A, 2001, vol. 43,
no. 10, p. 1814.
102. Tonoyan, A.O., Arakelova, E.R., Airapetyan, S.M., Mamalis, A.G., and Davtyan, S.P., Arm. Khim. Zh., 2001,
vol. 54, nos. 1–2, p. 65.
103. Airapetyan, S.M., Tonoyan, A.O., Arakelova, E.R., Saakyan, A.A., and Davtyan, S.P., Arm. Khim. Zh., 2001,
vol. 52, nos. 3–4, p. 45.
104. Tonoyan, A.O., Airapetyan, S.M., and Davtyan, S.P., Arm. Khim. Zh., 2001, vol. 52, nos. 3–4, p. 76.
105. Tonoyan, A.O., Davtyan, S.P., Airapetyan, S.M., Arakelova, E.R., and Saakyan, A.A., Arm Patent
NP20000017, 2001.
106. Ayrapetyan, S.M., Tonoyan, A.O., Kirakosyan, N.N., Chachatryan, A.R., and Davtyan, S.P., About Possibility
of Intercalation during Formation of Superconducting Polymer–Ceramic Compositions, Enicolopov’s Read
ings: Within the Framework of Int. Sci. Conf. of SEUA, Yerevan, 2003, p. 19.
107. Davtyan, S.P., Tonoyan, A.O., Hayrapetyan, S.M., and Manukyan, L.S., J. Mater. Process. Technol., 2005,
vol. 160, no. 3, p. 306.
108. Davtyan, S.P., Tonoyan, A.O., and Schick, C., Compos. Interfaces, 2006, vol. 13, nos. 4–6, p. 535.
109. Davtyan, S.P. and Tonoyan, A.O., in Chemistry of Advanced Compounds and Materials, Lekishvili N. and Zaikov
G.E., Eds., Nova Science, 2008, p. 3.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 49

110. Davtyan, S.P., Tonoyan, A.O., Schick, C., and Sargcyan, A.G., J. Mater. Process. Technol., 2007, vol. 163,
no. 5, p. 734.
111. Davtyan, S.P. and Tonoyan, A.O., Vysokotemperaturnye sverkhprovodniki. Sverkhprovodyashchie polimer
keramicheskie nanokompozity (HighTc Superconductors and Superconducting Polymer–Ceramic Nanocom
posites), Yerevan: Limush, 2008.
112. Davtyan, S.P., Tonoyan, A.O., and Schick, C., Materials, 2009, vol. 2, p. 2154.
113. Davtyan, S.P., Berlin, A.A., and Tonoyan, A.O., Obz. Zh. Khim., 2011, vol. 1, no. 1, p. 58.
114. Privalko, V.P. and Titov, G.V., Vysokomol. Soedin., Ser. A, 1978, vol. 21, p. 380.
115. Liu, X. and Wu, Q., Polymer, 2001, vol. 42, p. 10 013.
116. Huang, X. and Brittain, W.J., Macromolecules, 2001, vol. 34, no. 10, p. 3255.
117. Lipatov, Ya.S. and Privalko, V.P., Vysokomol. Soedin., Ser. B, 1972, vol. 14, no. 7, p. 1843.
118. Kalogeras, I.M. and Neagu, N.R., Eur. Phys. J. E, 2004, vol. 14, p. 193.
119. Bershtein, V.A., Egorova, L.M., Yakushev, P.N., Pissis, P., Sysel, P., and Brozova, L., J. Polym. Sci., Part B:
Polym. Phys., 2002, vol. 40, p. 1056.
120. Tabtiang, A., Lumlong. S., and Venables, R.A., Eur. Polym. J., 2000, vol. 36, p. 2559.
121. Arrighi, V., McEwena. I.J., Qiana, H., Serrano Prieto, M.B., Polymer, 2003, vol. 44, p. 6259.
122. Ash, B.J., Schadler, L.S., and Siegel, R.W., Mater. Lett., 2002, vol. 55, p. 83.
123. Ash, B.J., Siegel, R.W., and Schadler, L.R., J. Polym. Sci., Part B: Polym. Phys., 2004, vol. 42, p. 4371.
124. Miwa, Y., Drews, A.R., and Schlick, S., Macromolecules, 2006, vol. 39, p. 3304.
125. Mijovic, J., Lee, H.K., Kenny, J., and Mays, J., Macromolecules, 2006, vol. 39, p. 2172.
126. Giannelis, E.P., Krishnamoorti, R., and Manias, E., Adv. Polym. Sci., 1999, vol. 138, p. 107.
127. Lee, D.C. and Jang, L.W., J. Appl. Polym. Sci., 1996, vol. 61, p. 1117.
128. Privalko, V.P., Lipatov, Ya.S., and Kercha, Ya.Ya., Vysokomol. Soedin., Ser. A, 1970, vol. 12, no. 6, p. 1520.
129. Mammeri, F., Rozes, L., Le Bourhis, E., and Sanchez, C., J. Eur. Ceram. Soc., 2006, vol. 26, p. 267.
130. Vieweg, S., Unger, R., Hempel, E., and Donth, E., J. NonCryst. Solids, 1998, vol. 235, p. 470.
131. Kirst, K.U., Kremer, F., and Litvinov, V.M., Macromolecules, 1993, vol. 26, p. 975.
132. Litvinov, V.M. and Spies, H.W., Makromol. Chem., 1991, vol. 192, p. 3005.
133. Lin, W.Y. and Blum, F.D., Macromolecules, 1998, vol. 31, p. 4135.
134. Yagrarov, M.S., Ionkin, V.S., and Gizatullina, Z.G., Vysokomol. Soedin.: Ser. A, 1969, vol. 11, no. 10, p. 2626.
135. Haraguchi, K. and Li, H.J., Macromolecules, 2006, vol. 39, p. 1898.
136. Tsagaropoulos, G. and Eisenberg, A., Macromolecules, 1995, vol. 28, p. 6067.
137. Tsagaropoulos, G. and Eisenberg, A., Macromolecules, 1995, vol. 28, p. 396.
138. Fragiadakis, D., Pissis, P., and Bokobza, L., Polymer, 2005, vol. 46, p. 6001.
139. Salaniwal, S., Kumar, S.K., and Douglas, J.F., Phys. Rev. Lett., 2002, vol. 8, no. 25, p. 258 301.
140. Donth, E., Glass Transition, Berlin: Springer, 2001.
141. Schick, C., Sukhorukov, D., and Schonhals, A., Macromol. Chem. Phys., 2001, vol. 202, no. 8, p. 1398.
142. Huth, H., Wang, L.M., Schick, C., and Richert, R., J. Chem. Phys., 2007, vol. 126, p. 104 503.
143. Schick, C., Dobbertin, J., Potter, M., Dehne, H., Hensel, A., Wurm, A., et al., J. Therm. Anal., 1997, vol. 49,
no. 1, p. 499.
144. Dobbertin, J., Hensel, A., and Schick, C., J. Therm. Anal., 1996, vol. 47, no. 4, p. 1027.
145. Xia, H. and Song, M., Thermochim. Acta, 2005, vol. 429, p. 1.
146. Klonos, P., Panagopoulou, A., Bokobza, L., Kyritsis, A., Peoglos, V., and Pissis, P., Polymer, 2010, vol. 51,
p. 5490.
147. Klonos, P., Panagopoulou, A., Kyritsis, A., Bokobza, L., and Pissis, P., J. NonCryst. Solids, 2011, vol. 357,
p. 610.
148. Suzuki, H., Grebowicz, J., and Wunderlich, B., Br. Polym. J., 1985, vol. 17, no. 1, p. 1.
149. Wunderlich, B., Prog. Polym. Sci., 2003, vol. 28, no. 3, p. 383.
150. Xu, H. and Cebe, P., Macromolecules, 2004, vol. 37, p. 2797.
151. Minakov, A.A., Mordvintsev, D.A., Tol, R., and Schick, C., Thermochim. Acta, 2006, vol. 442, p. 25.
152. Minakov, A.A., Mordvintsev, D.A., and Schick, C., Polymer, 2004, vol. 45, no. 11, p. 3755.
153. Ishida, Y., Yamafuji, K., Ito, H., and Takayanagi, M., Kolloid Z. Z. Polym., 1962, vol. 184, no. 2, p. 97.
154. Donth, E., J. NonCryst. Solids, 1982, vol. 53, p. 325.
155. Schick, C. and Donth, E., Phys. Scr., 1991, vol. 43, no. 4, p. 423.
156. Schubnell, M. and Schawe, J.E.K., Int. J. Pharm., 2001, vol. 217, nos. 1–2, p. 173.
157. Wunderlich, B., Pure Appl. Chem., 1995, vol. 67, no. 6, p. 1019.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


50 DAVTYAN et al.

158. Dejong, B.H.W.S., Slaats, P.G.G., Super, H.T.J., Veldman, N., and Spek, A.L., J. NonCryst. Solids, 1994,
vol. 176, p. 164.
159. Hodge, I.M., J. NonCryst. Solids, 1994, vol. 169, p. 211.
160. Hay, J.N., Pure Appl. Chem., 1995, vol. 67, no. 11, p. 1855.
161. Lu, H. and Nutt, S., Macromolecules, 2003, vol. 36, p. 4010.
162. Lu, H. and Nutt, S., Macromol. Chem. Phys., 2003, vol. 204, p. 1832.
163. Adam, G. and Gibbs, J.H., J. Chem. Phys., 1965, vol. 43, p. 139.
164. Scheidler, P., Kob, W., and Binder, K., Europhys. Lett., 2002, vol. 59, no. 5, p. 701.
165. Tonoyan, A.O., Bagdasaryan, A.E., Manukyan, L.S., Kirakosyan, N.N., and Davtyan, S.P., Izv. Nats. Aakd.
Nauk Arm. Gos. Inzh. Univ. Arm., 2003, vol. 56, no. 2, p. 20.
166. Davtyan, S.P., Tonoyan, A.O., Manukyan, L.S., and Ayrapetyan, S.M., Eur. Polym. J., 2002, vol. 12, p. 2423.
167. Davtyan, D.S., Bagdasaryan, A.E., Tonoyan, A.O., and Davtyan, S.P., Khim. Fiz., 2000, vol. 19, no. 9, p. 83.
168. Davtyan, D.S., Bagdasaryan, A.E., Tonoyan, A.O., and Davtyan, S.P., Khim. Fiz., 2000, vol. 19, no. 9, p. 100.
169. López, E., Márquez, A., Flores, S., Ibarra, R., Hernández, C., Yacamán, M., and Zaragoza, A., Proc. World
Polymer Congr.–Macro 2006, 41st Int. Symp. on Macromolecules, Rio de Janeiro, 2006.
170. Davtyan, S.P., Zhirkov, V.P., and Vol’fson, S.A., Usp. Khim., 1984, vol. 53, p. 251.
171. Khanukaev, B.B., Kozhushner, M.A., and Enikolopyan, N.S., Fiz. Goreniya Vzryva, 1974, vol. 10, no. 1, p. 22.
172. Khanukaev, B.B., Kozhushner, M.A., and Enikolopyan, N.S., Fiz. Goreniya Vzryva, 1974, vol. 10, no. 5, p. 643.
173. Davtyan, S.P., Surkov, N.F., Davtyan, S.P., Rozenberg, B.A., and Enikolopyan, N.S., Dokl. Akad. Nauk, 1977,
vol. 232, p. 64.
174. Chechilo, N.M. and Enikolopyan, N.S., Dokl. Akad. Nauk, 1975, vol. 221, p. 1140.
175. Chechilo, N.M. and Enikolopyan, N.S., Dokl. Akad. Nauk, 1976, vol. 230, p. 160.
176. Pojman, J.A., Willis, J., Fortenberry, D., Ilyashenko, V., and Khan, A.M., J. Polym. Sci., Part A: Polym. Chem.,
1995, vol. 33, p. 643.
177. Manukyan, L.S., Airapetyan, S.M., Tonoyan, A.O., and Davtyan, S.P., Izv. Nats. Akad. Nauk Arm. Gos. Inzh.
Univ. Arm., 2003, vol. 56, no. 1, p. 52.
178. Yangchuan Ke, Chenfen Long, and Zongneng Qi, J. Appl. Polym. Sci., 1999, vol. 71, p. 1139.
179. Kornmann, X., Lindberg, H., and Berglund, L.A., Polymer, 2001, vol. 42, p. 4493.
180. Gopakumar, T.G., Lee, J.A., Kontopoulou, M., and Parent, J.S., Polymer, 2002, vol. 43, p. 5483.
181. Joo Young Nam, Suprakas Sinha Ray, and Masami Okamoto, Macromolecules, 2003, vol. 36, p. 7126.
182. Lincoln, D.M. and Vaia, R.A., Macromolecules, 2004, vol. 37, p. 4554.
183. Amerio, E., Sangermano, M., Malucelli, G., Priola, A., and Voit, B., Polymer, 2005, vol. 46, p. 11 241.
184. Román, F., Montserrat, S., and Hutchinson, J.M., J. Thermal Anal. Calorim., 2007, vol. 87, no. 1, p. 113.
185. Rosso, P. and Lin Ye, Macromol. Rapid Commun., 2007, vol. 28, p. 121.
186. Dae Su Kim, JinTae Kim, and Won Bum Woo, J. Appl. Polym. Sci., 2005, vol. 96, p. 1641.
187. Benson, R. and Cairos, T., J. Am. Chem. Soc., 1948, vol. 70, no. 5, p. 2115.
188. Volkova, T.M., Kinetics of Activated @Caprolactam Anion Polymerization, Extended Abstract of Cand. Sci.
(Chem.) Dissertation, Moscow, 1982.
189. Davtyan, S.P., Neizotermicheskie metody sinteza polimerov. Teoriya i praktika protsessov adiabaticheskoi polimer
izatsii (Nonisothermal Polymer Synthesis Processes: Theory and Practice of Adiabatic Polymerization Pro
cesses), Yerevan: Asogik, 2004.
190. Begishev, V.P., Bolgov, S.A., Malkin, A.Ya., Subbotina, N.I., and Frolov, V.G., Vysokomol. Soedin., Ser. B,
1980, vol. 22, no. 2, p 124.
191. Davtyan, S.P., Schick, C., Tonoyan, A.O., and Zohrabyan, D., ePolymers, 2011, no. 040.
192. Litmanovich, A.A. and Papisov, I.M., Vysokomol. Soedin., Ser. B, 1997, vol. 39, p. 323.
193. Litmanovich, A.A., Bogdanov, A.G., and Papisov, I.M., Vysokomol. Soedin., Ser. B, 1997, vol. 43, p. 2023.
194. Litmanovich, O.E., Litmanovich, A.A., and Papisov, I.M., Vysokomol. Soedin., Ser. B, 2000, vol. 42, p. 670.
195. Litmanovich, A.A., Bogdanov, A.G., and Papisov, I.M., Vysokomol. Soedin., Ser. B, 2001, vol. 43, p. 135.
196. Litmanovich, O.E., Vysokomol. Soedin., Ser. A, 2008, vol. 50, p. 1370.
197. Litmanovich, O.E. and Litmanovich, A.A., Vysokomol. Soedin. Ser. A, 2007, vol. 4, p. 674.
198. Litmanovich, O.E., Litmanovich, A.A., and Papisov, I.M., Vysokomol. Soedin., Ser. A, 2007, vol. 49, p. 684.
199. Gole, A. and Murphy, C.J., Langmuir, 2005, vol. 21, p. 10 756.
200. Rudin, A., Schreiber, H.P., and Waldman, M.H., Ind. Eng. Chem., 1961, vol. 53, p. 137.
201. Torsueva, E.S. and Shlyapnikov, Yu.A., Vysokomol. Soedin., Ser. A, 1973, vol. 175, p. 637.
202. Afanasiadi, L.I., Blank, A.B., Rvichko, L.A., Rotok, L.A., Litvinenko, Yu.G., Moghilko, E.T., Nastova, Z.M.,
Pavlyik, V.A., Pirogov, A.S., Pulyayeva, I.V., and Sheshina, S.G., Bull. Mater. Sci., 1991, vol. 14, no. 2, p. 335.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013


SYNTHESIS AND PROPERTIES OF PARTICLEFILLED 51

203. Rybkina, G.G., Ryabin, V.A., Moruni, M.S., Naberezhneva, E.L., and Vostretsova, A.B., Sverkhprovodimost:
Fiz., Khim., Tekh., 1993, vol. 6, nos. 11–12, p. 2126.
204. Krasil’nikov, V.N., Aptsigina, V.V., and Bezuev, G.V., Sverkhprovodimost: Fiz., Khim., Tekh., 1993, vol. 7, no. 1,
p. 135.
205. Enikolopyan, N.S. and Vol’fson, S.A., Khimiya i tekhnologiya poliformal’degida (Chemistry and Technology of
Polyformaldehyde), Moscow: Khimiya, 1968.
206. Madorsky, S.L., Thermal Decomposition of Organic Polymers, New York: Wiley–Interscience, 1964.
207. Wfng Huagin, Zhang Shiyian, Jin Tongheng, and Han Shiyuan, Mod. Phys. Lett. B, 1987, vol. 1, no. 8, p. 289.
208. Kuz’micheva, G.M. and Khlybov, E.P., Neorg. Mater., 1990, vol. 26, no. 6, p. 1264.
209. Goodentough, Y.B., Demazen, G., Pouchard, M., and Hagenmuller, P., J. Solid State Chem., 1973, vol. 8,
p. 325.
210. Ganduly, P. and Rao, C.N.R., Mater. Res. Bull., 1973, vol. 8, no. 4, p. 405.
211. Rao, C.N.R., Parkash, O., and Ganguly, P., J. Solid State Chem., 1975, vol. 15, no. 2, p. 186.
212. Mirzoyan, A.A., Garibyan, T.A., Gazaryan, K.G., and Arutyunyan, M.G., Sverkhprovodimost: Fiz., Khim.,
Tekh., 1991, vol. 4, no. 9, p. 986.
213. Ivanov, S.S. and Dmitrienko, A.B., Usp. Khim., 1982, vol. 51, no. 7, p. 1178.
214. Bruk, M.A., Pavlov, S.A., and Apkin, A.D., Radiat. Phys. Chem., 1981, vol. 17, no. 2, p. 113.
215. Farris, R.J., J. Appl. Polym. Sci., 1964, vol. 8, p. 25.
216. Topolkaraev, V.A., Tovmasyan, Yu.M., Dubnikova, I.L., et al., Dokl. Akad. Nauk SSSR, 1986, vol. 290, no. 6,
p. 1418.
217. Bartenev, G.M. and Zuev, Yu.S., Prochnost’ i razrushenie vysokoelastichnykh materialov (Strength and Fracture
of Highly Elastic Materials), Moscow, 1964, p. 195.
218. Rheology: Theory and Applications, Eirich, F.R., Ed., New York: Academic, 1956.
219. Yim, A. and Pierre, E.St., J. Polym. Sci., Part B: Polym. Phys. 1969, vol. 7, p. 237.
220. Braus, G. and Gruver, J.T., J. Polym. Sci., Part A: Polym Chem., 1970, vol. 2, p. 571.
221. Ball, G.L. and Salyer, I.O., J. Acoust. Soc. Am., 1966, vol. 39, p. 663.
222. Boehme, R.D., J. Appl. Polym. Sci., 1968, vol. 12, p. 1097.
223. Thurn, H., Kunststoffe, 1960, vol. 50, p. 606.
224. Romanyukha, A.A., Shvachko, Yu.N., and Ustinov, V.V., Usp. Fiz. Nauk, 1991, vol. 161, p. 37.
225. Asaturyan, R.A., Mod. Phys. Lett. B, 1993, vol. 7, p. 2043.
226. Asaturyan, R.A., Sarkisyan, D.A., Ignatyan, E.H., and Begoian, K.G., Solid State Commun., 1995, vol. 95,
p. 389.
227. Nicolais, L. and Narkis, M., Polym. Eng. Sci., 1971, vol. 11, p. 194.
228. Clarne, R. and Uher, C., Adv. Phys., 1984, vol. 33, p. 469.
229. Beille, J., Cabanel, R., Chaillaut, C., Chevaler, B., et al., Acad. Sci., Ser, 1987, vol. 304, p. 1094.
230. Bukovski, Z., Horin, R., and Rogacki, K., J. LessCommon Met., 1988, vol. 144, p. 153.
231. Enz, C.P., Helv. Phys. Acta, 1988, vol. 61, p. 741.
232. Rao, C.N.R., Phys. C (Amsterdam, Neth.), 1988, vols. 153–155, p. 1762.
233. Balzarotti, A., DeCrescenzi, M., Motta, N., Patella, F., and Scarlata, A., Phys. Rev. B: Condens. Matter Mater.
Phys., 1988, vol. 38, p. 6461.
234. Jorgesen, J.D., Veal, B.W., Paulikas, A.P., Nowicki, L.J., et al., L, Phys. Rev. B: Condens. Matter Mater. Phys.,
1990, vol. 41, p. 1863.
235. Goyot, H., Surf. Sci., 1992, nos. 269–270, p. 1082.
236. Chemistry of Oxide Superconductors, Rao, C.N.R., Ed., Oxford: Blackwell Scientific, 1988.
237. Baikov, Yu.M., Shalkova, E.K., and Ushakova, T.A., Sverkhprovodimost: Fiz., Khim., Tekh., 1993, vol. 6, p. 449.
238. Shapligin, I.S., Kakhap, B.G., and Lazarev, V.B., J. Inorg. Chem., 1979, vol. 24, p. 1478.

REVIEW JOURNAL OF CHEMISTRY Vol. 3 No. 1 2013

SPELL: 1. deagglomeration, 2. enthalpy, 3. semicrystalline, 4. vinylcapropropyllactam, 5.


orthorhombic

View publication stats

Potrebbero piacerti anche