You are on page 1of 6

Polymer Blends: Structure and Properties☆

Sam Rostami, ICI Plc, Middlesbrough, United Kingdom


Syarifah N Aqida, Universiti Malaysia Pahang, Pekan, Pahang, Malaysia
r 2018 Elsevier Inc. All rights reserved.

1 Introduction 1
2 Immiscible Blends 1
3 Miscible Blends 3
4 Conclusions 5
References 5
Further Reading 6

1 Introduction

The science and technology of polymeric mixtures has attracted much academic and industrial interest over the last few decades. As
a result, the basic rules of blending are now well established and numerous advances have been made to underpin the role of
molecular interactions on the phase behavior of blends made from high molecular weight polymers.
In homogeneous blends, spectroscopy methods, such as FTIR and NMR, are now commonly used to study the type and nature
of molecular interactions, whereas advanced thermodynamic models are employed to model the phase behavior and other
physical properties of the mixtures. Results show that two different chains of relatively high molecular weight polymers are only
intimately mixed when their favorable intermolecular interactions are accompanied by a large and negative change in their
enthalpy of mixing. In the absence of any strong intermolecular interaction, the gain in the entropy of mixing, chain conformation,
and its variation with an applied field become the dominant factors in determining the homogeneity or heterogeneity of polymeric
mixtures. On this basis, one may conclude that Lewis base polymers (such as poly(methyl methacrylate)) may mix with Lewis acid
polymers (such as poly(vinyl chloride)).Mixing of two nonpolar polymers such as polypropylene and polyethylene is greatly
influenced by the presence of short-ranged structures and their chemical moieties (Choi, 2000). Furthermore, polar and nonpolar
polymers should not intimately mix. The above analysis is only true as a first approximation. However, there are other factors, such
as the change in free volume, isothermal compressibility, thermal expansion, etc., that affect the phase behavior of polymeric
mixtures (Eichinger and Flory, 1968; Olabisi and Shaw, 1979; Sanchez et al., 1994). More details of theoretical considerations and
various thermodynamic models that are being developed and used to explore the phase separation phenomena in polymer blends
are given elsewhere (Miles and Rostami, 1992; Sandler, 1994) and will not be described in any depth here. Suffice to say that the
majority of the existing models rely on some experimental data to ‘‘train’’ the models and then simulate the phase behavior of
polymeric mixtures.
Emerging approaches, such as the group contribution (GC) estimation (Fredenslund and Sorensen, 1994) and polymer
reference interaction site models (Honeycutt, 1994; Case and Honeycutt, 1994), can potentially predict the phase behavior
of polymeric mixtures without the need for an experimental input by using Bagley miscibility maps. The plot in Fig. 1 have been
developed using solubility parameters of polymer/polymer blends with poly(ethylene oxide) (PEO) and 55 blending
partners through GC method that was proposed by Hoy, Hoftyzer-Van Krevelen, Stefanis-Panayiotou and Yamamoto (Meaurio
et al., 2017).
Absence of a strong and favorable interaction can easily lead to a phase separation in polymer blends. In other words, when the
change in free energy of mixing between two components is positive, or more appropriately when the second derivative of change
in free energy in respect of the composition is negative, the phase separation prevails. This is the case in the majority of polymeric
mixtures that have been studied.
In this article, the formation of phase structures in miscible and immiscible blends is considered and advantages and dis-
advantages of each structure are highlighted with the support of some experimental data. The term ‘‘miscible’’ is used to describe
those blends that are homogeneous at a molecular level. Blends that are homogeneous at some temperatures and phase separate at
other accessible temperature regions are referred to as partially or near-miscible blends. The term ‘‘immiscible’’ is used to describe
phase-separated morphologies.
From a commercial point of view, polymer blends (miscible, partially miscible, and immiscible) have enjoyed a steady annual
growth rate of 6%–7% since the early 1980s. The majority of blends used in commercial applications are by and large of
immiscible nature. They have found niche applications in commodity, engineering, and tailor-made products. Utracki (1998)
gives a survey of commercially developed polymer blends.


Change History: February 2018. Syarifah N Aqida added abstract; keywords; expanded text with additional review materials; and added list of reference.

Reference Module in Materials Science and Materials Engineering doi:10.1016/B978-0-12-803581-8.11301-3 1


2 Polymer Blends: Structure and Properties

Fig. 1 Bagley plot for polymer blends with PEO calculated using the GC method (●), weak interacting miscible polymers (▲), strong hydrogen
bonding miscible polymers (♦), weak interacting non miscible polymers (■), strong hydrogen bonding non-miscible polymers. Reproduced from
Meaurio, E., Sanchez-Rexach, E., Zuza, E., et al., 2017. Predicting miscibility in polymer blends using the Bagley plot: Blends with poly(ethylene
oxide). Polymer 113, 295–309.

2 Immiscible Blends

Strong and favorable molecular interactions between two different polymers are rare and the gain in entropy of mixing becomes
increasingly small as the molecular weight of each polymer increases. Most polymeric mixtures are therefore expected to phase separate on
mixing. The degree of their phase separation is governed by the interfacial energy between the two components. The tendency for a gross
phase separation is higher when the interfacial energy is higher and vice versa. It is difficult to predict reliably the interfacial energy between
two polymers. Similarly, direct experimental measurement of the interfacial energy between two high molecular weight polymer chains is
tedious and often unreliable (Luciani et al., 1997). Direct methods such as breaking thread, sessile drop, or contact angle techniques
(Wu, 1982) developed for measuring the interfacial energy between low molecular weight materials are not readily adaptable for high
molecular weight polymers. Indirect methods, such as surface reflectivity (Richards and Peace, 1999; Higgins and Benoit, 1994) are often
used to measure the interfacial width and calculate the interfacial energy between two high molecular weight polymers.
Practical experiences, however, show that polymer blends with an interfacial energy greater than 5 mNm 1 are grossly phase
separated with phase sizes of well over 100 nm in diameter, depending on the viscosities of the two components and the mixing
method used. The grossly phase-separated morphology is generally unsuitable for many practical applications owing to its adverse
effects on the physical and mechanical properties of the polymeric mixtures. In order to be able to use such a system, their interfacial
energy has to be reduced to, ideally, below 2 mNm 1. This is often achieved by an addition of a third component generally known as a
‘‘compatibilizer’’ or a change of functional groups in the polymer chains. Theoretical and experimental studies (Shull, 1991; Bucknall
et al., 1992) show that compatibilizers predominantly reside at the interfacial regions between the two components in order to reduce
their interfacial energy. In such a case, the effective compatibilizer should reduce the phase sizes and enhance the fracture toughness of
the blends. To date, researchers have been introducing nanoparticles in immiscible polymer blend for tailored morphology, phase
diagram and properties (Liu et al., 2017). In Fig. 2, the number of publications on the morphology of nanoparticles filled polymer
blends enhanced within the last decade as reviewed by de Luna and Filippone (2016).
An example of compatibilizer addition is in nylon 66 and polypropylene (PP). The interfacial energy between nylon 66 and PP
measured at 2851C is found to be about 15.5 mNm 1 (Miles and Rostami, 1992). A 1:1 ratio of PP:nylon 66, melt blended in a twin-
screw extruder, produces a blend with mean particle size of 500 nm of PP dispersed in nylon 66. Furthermore, it is noticed that the
phase structure changes from the surface to the center of the specimens produced by an injection molding operation. This is
presumably due to differential shear stress and the flow experienced during the injection molding operation by such a large phase-
separated morphology. This type of morphology is therefore sensitive to the shear rate and changes variably when processed in
different equipment. Owing to the gross phase separation between the two components, the measured fracture toughness of the blend
is also poor, as shown in Table 1. However, addition of a maleated polypropylene (MPP) as a compatibilizing agent not only reduces
the PP domains but also enhances the fracture energy of the mixture. In this case the maleic anhydride groups, randomly grafted onto
PP chains, are expected to react with amine end groups and/or the amide functional groups in nylon 66. The PP components of MPP
should reside in the PP phase thus helping to reduce the interfacial repulsive forces between PP and nylon 66.
Polymer Blends: Structure and Properties 3

Fig. 2 Evolution of number of publications on the morphology of unfilled (left axis) and nanoparticle-filled (right axis) polymer blends.
Reproduced from de Luna, M.S., Filippone, G., 2016. Effects of nanoparticles on the morphology of immiscible polymer blends – Challenges and
opportunities, European Polymer Journal 79, 198–218.

Table 1 Some mechanical properties of compatibilized and uncompatibilized blends of PP and nylon 66

Composition (PA:PP:MPP) (wt%) Modulus (MPa) Yield stress (MPa) GIc (kJ m 2) Average particle size (nm)

75:25:0 2.68 42.1 2.39 500


72.5:22.5:5 2.55 50.4 3.1 100
25:75:0 1.38 17.8 1.72 600
22.5:72.5:5 1.49 28.0 3.45 150

In practice, the migration of MPP to the interfaces may be controlled by a diffusion process that requires a long time to reach
‘‘equilibrium’’ and saturate the interfacial regions. To assist this process, a higher level of MPP, far more than is theoretically needed, is
added to both PP and nylon 66. As shown in Table 1, the average particle size of PP reduces and the fracture toughness of the blend
increases in the presence of the MPP. These are the two most affected properties observed in the presence of the compatibilizer.
In another example, melt blended mixtures of poly(ether sulfone) (PES) and nylon 66 with an estimated interfacial energy of
B2 mNm 1 were studied. The transmission electron microscopy of the blends made by a melt processing technique reveals fine
phase morphology between PES and nylon 66. For a 30:70 wt% PES in nylon 66, an almost uniform particle of PES, with a
diameter of around 2 nm, is dispersed in nylon 66. Most importantly the phase morphology remains fairly uniform throughout
the section of the specimens produced by the injection molding operation. The low interfacial energy between the two compo-
nents is thought to be responsible for the robust phase structure of the blend, which in turn ensures reproducible mechanical
properties. The reproducibility of mechanical properties of the PES: nylon 66 blends even when they are made at two different
locations and by using different sets of processing equipment is demonstrated in Table 2.
Since no copolymer of PES and nylon can be synthesized, other types of compatibilizing agents, such the phenoxy resin that is
partially miscible with PES and can chemically react with nylon 66, are used. As shown in Table 3, the addition of phenoxy resin
has no significant effect on the morphology or the toughness of the blend. Moreover, the morphology remains unchanged in the
presence of compatibilizer used. Although the phenoxy resin may not be an ideal compatibilizer, similar results are observed with
other compatibilizing agents that could possibly be used in this case.
Results in Tables 1, 2, and 3 suggest that blends with high interfacial energy and large phase sizes can benefit mostly from the
addition of a suitable compatibilizing agent whereas blends with low interfacial energy and small particle size may not require any
interfacial modification. In fact, experience of these blends and other similar systems shows that phase morphologies with phase
sizes below 5 nm are generally robust enough for commercial applications and may not require compatibilization.

3 Miscible Blends

As mentioned above, homogeneous blends are mainly formed as a result of strong and favorable interactions, such as hydrogen
bonding, between two different chains of high molecular weight polymers. The interactions are normally frozen-in below the glass
transition temperature (and crystalline melting temperature when semicrystalline components are involved) of the mixture and
grow weaker as the temperature increases. The stronger the interaction the higher the kinetic energy, in the form of heat, is required
to separate the two chains. At some high temperatures the molecular interactions between the two chains dissociate to cause phase
separation of a type known as the lower critical solution temperature.
4 Polymer Blends: Structure and Properties

Table 2 Some mechanical properties of PES:nylon 66 blends made in the UK and Japan

Composition (PES:nylon 66) (wt%) Tensile strength (MPa) Flexural modulus (GPa) Notched izod impact strength (kJ m 2)

UK Japan UK Japan UK Japan

30:70a 74 80 2.7 2.65 6 7


30:70b 78 83 2.8 2.6 5.2 6.7
a
Contains a high molecular weight nylon 66.
b
Contains standard molecular weight nylon 66.

Table 3 Some properties of PES:nylon 66 blends in the presence of phenoxy resin

Composition (PES:nylon 66:phenoxy) (wt%) Tensile yield strength (MPa) Tensile modulus (GPa) Fracture energy, GIc (kJ m 2)

30:70:0 74.6 2.69 1.34


28.5:68.5:3 67.7 2.65 1.1

Table 4 Tg and the level of crystallinity in PEK:PEI miscible blends

Composition (PEK:PEI) (wt%) Tag (1C) Crystallinitya (%) Methylene chloride uptakeb (%)

100:0 155 38 o1
80:20 167 28 o1
60:40 171 22 1.1
50:50 179 19 4
20:80 207 23
0:100 221
a 1
Data obtained using samples quenched from 4001C in a differential scanning calorimeter. A heating rate of 101C min is used for the reheat process.
b
Annealed samples for 1 h at 3001C.

The phase separation process is accompanied by a number of changes in physical properties of the mixture that are detected, for
example, by direct observations, scattering (Higgins and Benoit, 1994), or spectroscopic methods. The equilibrium and kinetics of
the phase separation process are the subject of other extensive experimental (Olabisi and Shaw, 1979; Ottenbrite et al., 1987) and
theoretical (Sandler, 1994) studies. However, limited work reported on the role of molecular interactions in the formation of
phase structures in polymeric mixtures. More importantly is a suitable theoretical model that is able to predict energy changes in
the system without the need for experimental data. To add to the complication, it was observed that under certain specific
conditions, and particularly when copolymers are involved, two polymers may become miscible at high temperatures and phase
separate on lowering the temperature. Recent advance in the modeling was using partial solvation parameters (PSP) based on
quantum mechanics and statistical thermodynamics to predict miscibility of polymer/polymer blend (Panayiotou, 2013). The
novel approach can determine melting point depression, fraction of hydrogen-bonded groups and critical composition of a
copolymer for miscibility with another polymer. Mazinani et al. (2017) and co-workers investigated the intermolecular interac-
tions and miscibility behavior of two commercial polyimide (PI) blend systems, Extem/Matrimid and Extem/U-Varnish at various
compositions. The Extem/U-Varnish system exhibits a single Tg in each composition that suggest miscibility behavior. The Tgs of
the polymer blends were also estimated by theoretical equations and compared with experimental data.
Experimentally, however, transparency and the glass transition temperature, Tg, are widely used to infer structural homogeneity
or heterogeneity in polymeric materials. When the two polymers are miscible their Tg lies between the Tg values of the parent
polymers. Every homogenous composition in this respect is expected to behave as a single-component polymer in its own right.
For example, poly(ether ketone) (PEK) forms miscible blends with poly(ether imide) (PEI) at all proportions. PEK is a semi-
crystalline polymer with a Tg of 1551C and an experimental crystalline melting temperature, Tm, of 3651C, whereas PEI is an
amorphous polymer with a Tg of 2201C. The addition of PEI to PEK raises the Tg of PEK that enhances its high-temperature usage
whereas PEK introduces crystallinity to PEI that enhances the environmental chemical resistance of PEI. Furthermore, the mis-
cibility in this case not only alters the Tg of the semicrystalline polymer but also has profound effects on the crystallization habit of
the crystalline regions in PEK. Consideration of this effect requires a separate article. However, most common aspects of the
crystallization of semicrystalline polymers in miscible blends are discussed elsewhere (Rostami, 1992, 2000). This topic will not be
further expanded here; instead results of some experimental work on miscible blends are presented.
Table 4 shows the Tg and percent crystallinity of PEK:PEI blends. These results are obtained using reheat runs in a differential
scanning calorimeter on samples quenched from 4001C. A heating rate of 101C min 1 is used for the reheat process.
Polymer Blends: Structure and Properties 5

Table 5 Properties of continuous carbon fiber-reinforced miscible blend

Property Miscible blends APC-2 as control

Flexural strength (GPa) 1.88 2.038


Flexural modulus (GPa) 129 134
Transverse flexural strength (MPa) 120 140
Short-beam shear strength (MPa) 100 105

PEK is a high-performance polymer suitable for many engineering applications. By raising its Tg and sacrificing some of its
crystallinity, PEK retains most of its high-temperature modulus at even higher temperatures that may make it suitable for new
applications. Furthermore, owing to the miscible nature of the blends they could also be reinforced, for example with continuous
carbon fibers, to produce high-performance engineering composites for applications that require both high-temperature and
strong chemical resistances. The balance between Tg, the level of crystallinity, and solvent uptake then becomes a critical issue. The
amounts of solvent uptakes, on annealed samples, after 24 h’ immersion in methylene chloride, are also shown in Table 4. As an
example, a 50:50 wt% blend of PEK with PEI that has a Tg of 1791C and B19 wt% crystallinity was reinforced with 60 wt% of
continuous carbon fiber in a solution prepreging operation. The prepregs were molded into test specimens using a hot press.
Results of mechanical testing on the composite with blend as its matrix are compared with those of ‘‘APC-2” in Table 5. APC-2 is
made of poly(ether ether ketone) reinforced with 63 wt% of continuous carbon fiber and used here as a control.
As shown in Table 5, good combinations of mechanical properties are obtained for the reinforced blend. Any other compo-
sition that is mentioned in Table 4 could similarly be treated depending on the application needs.

4 Conclusions

The structural formation and its effects on physical and mechanical properties of polymer blends are described. It is shown that the
phase structures of polymer blends made of high molecular weight materials are predominately influenced by the nature of the
molecular interactions between the two separate chains. When strong molecular interactions, such as hydrogen bonding, are
present between two polymers they may become miscible on mixing. When the interactions are weak or are unfavorable, the phase
separation prevails. In phase-separated blends, the extent of phase separation is influenced by the magnitude of the interfacial
energy between the two components. The relationship between interfacial energy and phase structures and the resulting properties
are explained. Experimental results are used to demonstrate these effects.
Progress made in theoretical predictions of miscibility or immiscibility between two components is highlighted. It is also
shown that advances in theoretical modeling and prediction plot are able to predict the phase structure of polymer blends without
experimental inputs. References are provided for a more in-depth follow-up of topics raised in this article.

References

Bucknall, D.G., Higgins, J.S., Rostami, S., 1992. The compatibilising effect of diblock copolymer on the morphology of immiscible blends. Polymer 33 (20), 4419.
Case, F.H., Honeycutt, J.D., 1994. Will my polymer mix? Trends Polym. Sci. 2 (8).
Choi, P., 2000. Polym. Commun. 41, 8741.
de Luna, M.S., Filippone, G., 2016. Effects of nanoparticles on the morphology of immiscible polymer blends – Challenges and opportunities. Eur. Polym. J. 79, 198–218.
Eichinger, B.E., Flory, P.J., 1968. Trans. Faraday Soc. 64, 2035.
Fredenslund, A., Sorensen, J.M., 1994. Group contribution estimation of phase equilibria I.S. Sandler. Models for Thermodynamic and Phase Equilibria Calculation. New York:
Marcel Dekker.
Higgins, J.S., Benoit, H.C., 1994. Polymer and Neutron Scattering. Oxford Science.
Honeycutt, J.D., 1994. Macromolecules 27 (19), 5377.
Liu, D., Li, W., Zhang, N., et al., 2017. Graphite oxide-driven miscibility in PVDF/PMMA blends: Assessment through dynamic rheology method. Eur. Polym. J. 96, 232–247.
Luciani, A., Champage, M.F., Utracki, L.A., 1997. Macromol. Symp. 126, 307.
Mazinani, S., Darvishmanesh, S., Ramazani, R., der Bruggen, B.V., 2017. Miscibility of polyimide blends: Physicochemical characterization of two high performance polyimide
polymers. React. Funct. Polym. 111, 88–101.
Meaurio E., Sanchez-Rexach E., Zuza E., et al., 2017. Predicting miscibility in polymer blends using the Bagley plot: Blends with poly(ethylene oxide). Polymer 113, 295-309.
Miles, I.S., Rostami, S., 1992. Multicomponent Polymer Systems. London: Chapman and Hall.
Olabisi, O., Shaw, M., 1979. Polymer–Polymer Miscibility. Academic Press.
Ottenbrite, R.M., Utraki, L.A., Inoue, S., 1987. Current Topics in Polymer Science. vol. II. New York: Macmillan.
Panayiotou, C., 2013. Polymer–polymer miscibility and partial solvation parameters. Polymer 54 (6), 1621–1638.
Richard, R.W., Peace, S.K., 1999. Polymer Surfaces and Interfaces III. Wiley.
Rostami, S., 1992. Crystallization behavior of semi-crystalline miscible blends. Polymer 31, 899.
Rostami, S., 2000. Advances in theory of crystalline melting point depression in miscible polymer blends. Eur. Polym. J 36, 2285.
Sanchez, I.C., Panayiotou, C.G., Sandler, I.S., 1994. Models for Thermodynamic and Phase Equilibria Calculation. New York: Marcel Dekker.
6 Polymer Blends: Structure and Properties

Sandler, I.S., 1994. Models for Thermodynamic and Phase Equilibria Calculation. New York: Marcel Dekker.
Shull, K.R., 1991. J. Chem. Phys. 94, 5723.
Utracki, L.A., 1998. Commercial Polymer Blends. Chapman and Hall.
Wu, S., 1982. Polymer Interface and Adhesion. New York: Marcel Dekker.

Further Reading
Shull, K.R., Kramer, E.J., 1990. Macromolecules 23, 4769.
Shull, K.R., Kramer, E.J., Hadzioannou, G., Tang, W., 1990. Macromolecules 23, 4780.

You might also like