Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Polymer Blends
Polymer Blends
Polymer Blends
Ebook782 pages6 hours

Polymer Blends

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Polymer Blends, Volume 2 aims to show the importance of mixed polymer systems as a major branch of macromolecular science and provides a broad background of principles and practices in this field. Starting from where the first volume left off, the book covers topics in the area of polymer blends in Chapters 11-23. Areas of coverage include interpenetrating polymer networks; interfacial agents for polymer blends; rubber modification of plastics; fracture phenomena; coextruded multilayer polymer films and sheets; polymeric plasticizers; and polyolefin blends and their applications. The book is recommended for scientists, technologists, and engineers in the academe, research, and related industry, especially those who wish to be updated with its advances as a science.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780323149778
Polymer Blends

Related to Polymer Blends

Related ebooks

Chemical Engineering For You

View More

Related articles

Reviews for Polymer Blends

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Polymer Blends - D.R. Paul

    Index

    Chapter 11

    Interpenetrating Polymer Networks

    D.A. Thomas and L.H. Sperling,     Materials Research Center, Coxe Laboratory, Lehigh University, Bethlehem, Pennsylvania

    Publisher Summary

    This chapter provides an overview of the interpenetrating polymer network (IPN) that denotes an entire class of materials, rather than a single molecular topology. An IPN is any material containing two polymers, each in network form. Both the polymers are synthesized or cross-linked in the presence of each other. It explains both simultaneous and sequential types of syntheses, and both yield distinguishable materials. Further, the chapter discusses the relationship among IPNs, blends, and grafts. A polymer blend is defined as a combination of two polymers without any chemical bonding between them, whereas a graft copolymer refers to a product prepared by the polymerization of monomer II in the presence of polymer I, with greater or lesser extent of actual graft copolymer formation. Recent electron microscope and kinetic evidence suggests that grafting in many systems is less extensive than previously believed, but it is still important. The chapter additionally highlights the morphology that indicates phase separation. The phases vary in amount, size, shape, sharpness of their interfaces, and the degree of continuity. The chapter also describes the factors that affect morphology of the material and the physical and mechanical behavior.

    I. Introduction

    A. Definition of an Interpenetrating Polymer Network

    B. Relationships among IPNs, Blends, and Grafts

    C. Historical Aspects

    D. IPN Nomenclature

    II. Isomeric Graft Copolymers and IPNs

    A. Group Theory Notation

    B. Rings and Binary Notation

    III. Survey of Synthetic Methods

    IV. Morphology

    A. Factors Affecting Morphology

    B. Review of Two-Phase Polymer Morphology

    V. Physical and Mechanical Behavior

    A. General Properties

    B. Glass Transitions and Viscoelastic Behavior

    C. Ultimate Behavior

    D. Recent Studies

    VI. Applications and Uses

    A. Existing Patents

    B. Potential Applications

    Acknowledgments

    References

    I. INTRODUCTION

    A. Definition of an Interpenetrating Polymer Network

    The term interpenetrating polymer network (IPN) denotes an entire class of materials, rather than a single molecular topology. In its broadest definition, an IPN is any material containing two polymers, each in network form. A practical restriction requires that the two polymers have been synthesized or cross-linked in the presence of each other. Both simultaneous and sequential types of syntheses have been explored; both yield distinguishable materials [1].

    One type of sequential IPN begins with the synthesis of cross-linked Polymer I. Monomer II, plus its own cross-linker and initiator, are swollen into Polymer I, and polymerized in situ [2]. Simultaneous interpenetrating networks (SINs) begin with a solution of both monomers and cross-linkers, which are then polymerized by noninterfering modes, such as addition and condensation reactions [3–5]. The two rates of polymerization, and their approach to gelation may be the same, or significantly different. Again, different products ensue. A third mode of IPN synthesis takes two latexes of linear polymers, mixes and coagulates them, and cross-links both components simultaneously. The product is called an interpenetrating elastomeric network (IEN) [6]. As is seen below, there are, in fact, many different ways that an IPN can be prepared; each yields a distinctive topology.

    The term IPN implies an interpenetration of the two polymer networks of some kind, and was coined before the full consequences of phase separation were realized [7]. Molecular interpenetration only occurs in the case of total mutual solubility; however, most IPNs phase separate to a greater or lesser extent. Given that the synthetic mode yields two networks, the extent of continuity of each network needs to be examined. If both networks are continuous throughout the sample, and the material is phase separated, the phases must interpenetrate in some way [8], and, thus, some IPN compositions are thought to contain two continuous phases.

    Thus, molecular interpenetration may be restricted or shared with supermolecular levels of interpenetration. True molecular interpenetration is thought to take place only at the phase boundaries in some cases. If the two polymers are chemically identical, the product is called a Millar IPN [7]. In this case, true compatibility is achieved, and the network chains are believed actually to interpenetrate at the molecular level (but see Section IV.A.5 for a possible reinterpretation).

    When only one of the polymers is cross-linked, the product is called a semi-IPN [9]. If the polymerizations are sequential in time, four semi-IPNs may be distinguished. If Polymer I is cross-linked and Polymer II is linear, the product is called a semi-IPN of the first kind, or semi-1 for short. If Polymer I is linear, and Polymer II cross-linked, a semi-2 results. The remaining two compositions are materialized by inverting the order of polymerization. Several recent reviews [10–13] have been written on the subject of IPNs.

    B. Relationships among IPNs, Blends, and Grafts

    A polymer blend may be defined as a combination of two polymers without any chemical bonding between them (Volume 1, Chapters 1 and 2). A graft copolymer refers to a product prepared by the polymerization of Monomer II in the intimate presence of Polymer I, with greater or lesser extent of actual graft copolymer formation. Recent electron microscope and kinetic evidence suggests that grafting in many systems is less extensive than previously believed, but still important. The graft copolymer may behave as a nonaqueous surface active agent, binding the two phases together at their interface (see Chapter 12). Several topologies of interest are illustrated schematically in Fig. 1. Since most IPNs involve polymerization of one polymer in the immediate presence of the other, they are also generally graft copolymers. They constitute a special class, however, since one or both polymers contain cross-links. The interesting properties of IPNs emerge when the deliberately introduced cross-links outnumber the accidentally introduced grafts. When this condition prevails, the cross-links dominate and control the morphology and hence influence most of the physical and mechanical behavior. However, the grafts are still present and usually contribute favorably to the behavior of the IPN as a material [14a,b].

    Fig. 1 Schematic diagram of some simple two-polymer combinations: (a) a polymer blend; (b) a graft copolymer; (c) a block copolymer; (d) a semi-IPN; (e) an IPN; (f) an AB cross-linked copolymer. The solid line represents Polymer I; dotted line represents Polymer II. Enlarged intersections represent cross-link sites.

    C. Historical Aspects

    The concept of an IPN goes back at least as far as 1941, when Staudinger and Hutchinson [15] first applied for a British patent. Their United States patent of 1951 describes the use of an IPN topology to prepare improved optically smooth plastic surfaces. The first use of the term interpenetrating polymer network was by Millar [7] in 1960. Frisch et al. [6] and Sperling [16] independently arrived at the IPN topology through different thought processes. Frisch conceived of the IPNs as the macro-molecular analog of the catenanes. The catenanes consist of interlocking ring structures, which are physically bound together like chain links. Sperling evolved IPNs as a method of producing finely divided phase domains without the need for mechanical mixing.

    D. IPN Nomenclature

    Some of the terminology used in this chapter is as follows:

    1. IPN: interpenetrating polymer network, the general term. Also used to indicate the time-sequential synthesis product.

    2. IEN: interpenetrating elastomeric network, originally used by Frisch et al. [6] to denote materials made by mixing two latexes, coagulating them, followed by independent cross-linking reactions.

    3. SIN: simultaneous interpenetrating network where both polymers are synthesized simultaneously, that is, by addition and condensation reactions.

    4. Semi-IPN: compositions, generally of one cross-linked polymer and one linear polymer.

    5. Semi-1, semi-2: semi-IPNs where, respectively, Polymer I or Polymer II is the cross-linked component.

    6. Interstitial composites: notation of Allen et al. [17] for materials which would be described as semi-SINs in the above notation system; both polymers are synthesized simultaneously, but one is linear, and another cross-linked.

    7. Gradient IPN: an IPN of nonuniform macroscopic composition, usually by nonequilibrium swelling in Monomer II, and polymerizing rapidly.

    A more complete nomenclature of this type was developed by Klempner and Frisch [18]. A more mathematically oriented nomenclature based on group and ring theory is presented below.

    II. ISOMERIC GRAFT COPOLYMERS AND IPNs

    An examination of the scientific and patent literature reveals that over 200 topologies of IPNs and related materials have been synthesized. By classical standards, most would be designated simply as graft copolymers, with little to distinguish one composition from another. In an attempt to correct the inadequacies of the nomenclature system, Sperling derived two types of systematic notation. The first is based on group theory concepts [19–21], and presents a detailed approach to the nomenclature of polymer blends, grafts, and IPNs. Later, a nomenclature system based on the ideas of mathematical rings was developed [22]. The latter is less detailed in nomenclature and offers improved insight into the interrelationships of the many materials. Even with simplifying assumptions, the two techniques show that more than 50,000 distinguishable combinations of the same two polymers are possible. The basics of the two systems are outlined in the following paragraphs.

    A. Group Theory Notation

    In terms of group theory concepts, the symbols Pi, Ci, and Gij represent the formation of the linear ith polymer, the cross-linked ith polymer, and the graft of the jth polymer onto the ith polymer, respectively, in the generalized multicomponent system. The symbols mi, ci, and gi stand for the monomer, cross-linker, and grafting agents, respectively. Parentheses indicate simultaneous processes, and brackets indicate materials synthesized separately, and later mixed and/or reacted. For example,

    (1)

    represents two linear polymers synthesized separately that were blended together and then both cross-linked [6].

    B. Rings and Binary Notation

    It has been assumed in the above that the system has some of the important characteristics of a mathematical group. Now, the subject will be cast in the form of a mathematical ring, as this leads to improved insight into the nature of the combinations and their proposed nomenclature. Two simplifications must be made:

    1. The use of reactive unit symbolism is being discarded.

    2. Materials that are slightly or accidentally grafted are considered to be blends for the present discussion. Only systems that are intentionally or extensively grafted are so considered.

    1. Thus,

    (2)

    means a blend of Polymers 1 and 2, reading from right to left.1 replaces the brackets used earlier, which stood for blending.

    The combinations involving chemical bonding between polymers is the multiplication, and its binary operation is designated o2. Thus,

    (3)

    indicates a cross-linked Polymer I, and

    (4)

    indicates the grafting of Polymer II onto a Polymer I backbone.

    The binary operation tables for o1 and o2 are shown in Tables I and II. The tables are infinite in size in the general case, with very complex structures possible. Two specific points should be noted:

    Table I

    Addition Table for Polymer Blendsa

    aIn the corresponding addition table for ordinary arithmetic, o1 means plus. Thus, 3 o1 5 = 8, 2 o1 3 = 5, etc., which can be set up in the same form as this polymer blend table. Ordinary addition and multiplication, in fact, form a ring, conforming to all its requirements.

    Table II

    Multiplication Table for Cross-Linked and Grafted Systems

    1. In the blend table, the commutative law of addition holds, and M12 = M21, where M stands for mixture or blend.

    2. Coefficients are omitted. In general, x1 yP1 = (x + y)P1, etc.

    However, simple combinations of the two tables are more illustrative. For example,

    (5)

    which makes use of the distributive property of multiplication over addition, and shows some of the characteristics of the notation. It should be noted that the accidentally or slightly grafted polymer combinations fit the blend table better than the cross-link-graft table. Thus, most of the IPNs synthesized to date may be considered analogous to a chemically induced blend of two cross-linked polymers (not a mechanical blend, certainly):

    (6)

    where I stands for an IPN.

    is a semi-IPN of polymers i and k, and h is 1 or 2, depending on whether the first or the second polymer so introduced contains the cross-links. It is obvious that much more complicated symbols could be evolved. For simplicity, however, combinations of existing symbols are employed, as in the right-hand side of Eq. (5).

    In Tables I and II, the elements in the left-hand column are employed first, and are operated on by the elements in the top row. This specifies the time order of the operations. Upon examination of the tables, a striking symmetry becomes apparent. Elements on either side of the diagonal from the upper left to the lower right are clearly related. For example, the qualitative relationship between G12 and G21 is obvious, but now we observe that they are quantitatively related in Table II, lying in symmetric positions across the diagonal. In fact, a function γ may be defined, which will cause the element to be moved to its corresponding position across the diagonal, and physically adopt its new structure. In general, there are two γ functions: γ1 for blends (Table I), and γ2 for grafts and cross-links (Table II). The transformation of element x into element y by the function γ is shown in Fig. 2. For example,

    Fig. 2 Illustrations on the application of γ, the function that calls for inverse materials within the cross-link graft table: (a) graft copolymers; (b) AB cross-linked polymers; (c) IPNs. The IPN case schematically shows the cellular structure of the second polymerized network.

    (7)

    Thus, γ1 transforms I12 to I21 within the blends, Table I. Since the more complex combinations lack common names, these are shown in Tables I and II as combinations of symbols.

    Also, more complex notations such as

    (8)

    are possible, representing the AB cross-linked copolymers. Equation (8) is read both vertically and horizontally, with time sequences moving from right to left. Quantities within the parenthesis are taken simultaneously. Equation (8) permits a fair (but still not entirely complete) representation of an AB-cross-linked copolymer [23], as distinguished from the simple graft copolymer analog, which does not constitute a network.†

    It should be emphasized that the above schemes do not fit all the requirements of a formal group or ring, but have sufficient similarities to warrant a comparison.

    III. SURVEY OF SYNTHETIC METHODS

    As explained above, the term IPN has gradually assumed a broader interpretation. Although the exact topologies possible run to many tens of thousands, only a relatively few laboratory techniques have been explored in depth.

    Sperling and Friedman [16] explored sequential IPNs, where Polymer Network I, poly(ethyl acrylate), was simultaneously photopolymerized and cross-linked with tetraethylene glycol dimethacrylate (TEGDM). Then portions of styrene monomer, cross-linker, and activator were swollen in, and allowed to equilibrate before a second photopolymerization was initiated (Fig. 3).

    Fig. 3 Schematic of the synthetic steps to make (a) IPNs and (b) SINs. Group theory notation is used. (From Touhsaent et al. [3].)

    Most of the sequential IPNs synthesized by Sperling et al. can be described through the group theory notation as

    (9)

    where the contents within parentheses indicate simultaneous processes. In terms of the ringlike notation, the sequential IPNs can be described as

    (10)

    Some variations on this theme have included latex IPNs [24], where each latex particle consists ideally of two molecules. In one set of experiments, however, linear styrene–butadiene rubber (SBR) was later cross-linked, then used for IPN formation with polystyrene (PS) [9, 25].

    If the cross-linked Polymer I is swollen in Monomer mix II for a limited period of time, and Polymerization II is carried out before equilibrium can be achieved, the product will contain a composition gradient. Gradient IPNs were prepared by Akovali et al. [26], Predecki [27], and Sperling and Thomas [28].

    An alternative synthetic technique involves the simultaneous polymerization of both polymers. The synthesis of SINs requires two independent and noninterfering reactions that can be carried out in the same reactor under the same conditions of heat, light, etc. Kim et al. [5] investigated simultaneous syntheses of polyurethanes (condensation polymerization) and polystyrene or poly(methyl methacrylate) (addition polymerization). Touhsaent et al. [3, 4] investigated the epoxy–poly(n-butyl acrylate) system, which also involves condensation and addition polymerization reactions run simultaneously. The group theory designation of the SIN formation reads

    (11)

    while the equivalent ringlike notation reads

    (12)

    Of course, both polymers need not be cross-linked. Many types of semi-IPNs exist. Donatelli et al. [9] synthesized both semi-1 and semi-2 compositions prepared from SBR elastomer and polystyrene, employing sequential polymerizations (see Fig. 1). Allen et al. [17] mixed urethane prepolymers with a triol curing agent, methyl methacrylate monomer, and activator. The ringlike designation for this synthesis reads

    (13)

    Allen designated his materials interstitial composites. A series of semi-SINs composed of polyurethanes and polyacrylates was recently reported by Yoon et al. [29].

    The AB cross-linked copolymers represent another interesting mode of joining two polymers. In this case, two polymers are required to make one network, as illustrated in Fig. If. Bamford et al. [23] have explored this method in some detail. Gardner and Baldwin [30] reported on the case of an AB cross-linked copolymer, where Polymer II was deliberately cross-linked. The graft copolymers prepared from epoxy compositions and CTBN elastomers [31] are recognized as a product having a topology much like that studied by Baldwin and Gardner [32]. In the latter case, the epoxy resin ordinarily cures to a densely cross-linked material. The carboxyl groups at the ends of the butadiene–nitrile rubber react with the epoxy to form the AB cross-linked analog. Another important case involves the castable polyesters [33]. In this last example, an unsaturated polyester is dissolved in styrene monomer, and on polymerization, the polystyrene grafts to and cross-links the polyester.

    IV. MORPHOLOGY

    Most IPNs and related materials investigated to date show phase separation. The phases, however, vary in amount, size, shape, sharpness of their interfaces, and degree of continuity. These aspects together constitute the morphology of the material, and the multitude of possible variations controls many of the material properties.

    Some aspects of morphology can be observed directly by transmission electron microscopy of stained and ultramicrotomed thin sections [34–36]. Unfortunately, the usual osmium tetroxide staining method is most successful only for polymers containing carbon–carbon double bonds, and many materials are not easily studied. Other aspects of morphology, such as phase continuity and interface characteristics, are best determined by combining information from chemical and mechanical tests with electron microscopy.

    A. Factors Affecting Morphology

    The factors that control the morphology of IPNs are now reasonably clear; they include chemical compatibility of the polymers, interfacial tension, cross-link densities of the networks, polymerization method, and the IPN composition. While these factors may be interrelated, they can often be varied independently. Their effects are summarized in this section.

    1. Compatibility

    A degree of compatibility between polymers is necessary for IPNs, because monomers or prepolymers must form solutions or swollen networks during synthesis.† Phase separation generally ensues in the course of polymerization, but the resulting phase domain size is smaller for higher-compatibility systems [38]. Huelck et al. [2] varied compatibility systematically in sequential IPNs, with poly(ethyl acrylate) (PEA) as Polymer I and copolymers of methyl methacrylate (MMA) and styrene (S) as Polymer II (Fig. 4). For PEA–PMMA, in which the components are isomeric and nearly compatible, dispersed phase domains less than 100 Å in size (fine structure) are found (Fig. 4a). This occurs because the high compatibility precludes phase separation until high conversion to PMMA. (An alternate explanation for the fine structure is given later.) With the much less compatible system PEA-PS (Fig. 4b), an additional cellular structure of about 1000 Å in size is found. Here, phase separation is thought to take place earlier in cellular regions rich in PS and S monomer. A second phase separation follows, leading to smaller dispersed PEA and PS domains in the cell walls.

    Fig. 4 Electron micrographs of IPNs of (a) 75/25 poly(ethyl acrylate)poly(methyl methacrylate), and (b) 50/50 poly(ethyl acrylate)/polystyrene. A small amount of butadiene was copolymerized with the poly(ethyl acrylate) to aid in osmium tetroxide staining. (From Huelck et al. [2]. Reprinted with permission from Macromolecules. Copyright by the American Chemical Society.)

    Styrene–butadiene rubber–polystyrene (SBR–PS) IPNs are relatively incompatible, even though they are both nonpolar polymers and their solubility parameters differ by only about 1.0. They show distinct phase separation and a cellular domain structure (Fig. 5) [39].

    Fig. 5 Electron micrographs of IPNs of 20/80 styrene–butadiene rubber/polystyrene. The rubber is cross-linked with (a) 0.1%, and (b) 0.2 % dicumyl peroxide. (From Donatelli et al. [39].)

    2. Cross-Linking

    Increased cross-link density in Polymer network I in an IPN clearly decreases the domain size of Polymer II [39]. This is illustrated by comparison of Figs. 5a and b. This effect is reasonable because the tighter initial network must restrict the size of the regions in which Polymer II can phase separate. The effect has also been rationalized by a semiempirical thermodynamic model [40], described later. Variation of cross-link density in the PS has little effect on the IPN morphology, indicating that the first network is controlling. In the extreme case of linear PS (forming a semi-1), the morphologies were somewhat less uniform, but cross-link density in the SBR had the same effect. With no cross-linking in the SBR, on the other hand, the SBR remained continuous but both phases were much coarser and irregular.

    Allen et al. [17] observed the same effect of the first network in polyurethane (PU)–PMMA materials, made by interstitial polymerization according to Eq. (13). The morphology shows roughly spherical domains of PMMA in a matrix of PU (Fig. 6). The PU network formed first from a solution of reactants including the MMA, but reaction conditions permitted varying tightness of the network. Although the PMMA was uncross-linked, its domain size varied from about 650 Å for tight PU networks to about 1800 Å for loose networks [41].

    Fig. 6 Electron micrograph of an interstitial polymer of 80/20 polyurethane/poly(methyl methacrylate). Poly(butadiene diols) in the polyurethane permit staining with osmium tetroxide. (From Allen et al. [41].)

    The AB cross-linked copolymers of Bamford et al. [23] are not strictly IPNs, since B (Polymer II, polychloroprene) chains are part of the same network as A [Polymer I, poly(vinyl trichloroacetate)] chains. However, like graft copolymers and IPNs, they do phase separate into distinct domains [42]. The Polymer II chains are seen to form the discontinuous phase, perhaps for two reasons: (1) similar to the sequential IPNs and nonphase inverted graft copolymers, Polymer II develops the less continuous phase morphology; and (2) Polymer II is also the minor component in this case.

    Meier [43, 44] considered the morphologies of diblock copolymers, and found that for dispersed spheres

    (14)

    where R represents the domain radius, k is a constant characteristic of the morphology (1.33 for spheres), C represents an experimental constant relating the unperturbed root-mean-square end-to-end distance of a chain to the square root of its molecular weight, and α is the ratio of the perturbed to unperturbed chain dimensions. From morphological dimensions, Eastmond and Smith [42]

    Enjoying the preview?
    Page 1 of 1